Sunteți pe pagina 1din 43

Grey cast iron[edit]

Main article: Grey iron


Grey cast iron is characterised by its graphitic microstructure, which causes fractures of the
material to have a grey appearance. It is the most commonly used cast iron and the most widely
used cast material based on weight. Most cast irons have a chemical composition of 2.54.0%
carbon, 13% silicon, and the remainder iron. Grey cast iron has lesstensile strength and shock
resistance than steel, but its compressive strength is comparable to low- and medium-carbon
steel.

White cast iron[edit]


White cast iron displays white fractured surface due to the presence of cementite. With a lower
silicon content (graphitizing agent) and faster cooling rate, the carbon in white cast
iron precipitates out of the melt as the metastable phase cementite, Fe3C, rather than graphite.
The cementite which precipitates from the melt forms as relatively large particles, usually in a
eutectic mixture, where the other phase is austenite (which on cooling might transform
to martensite). These eutectic carbides are much too large to provide precipitation hardening (as
in some steels, where cementite precipitates might inhibit plastic deformation by impeding the
movement of dislocations through the ferrite matrix). Rather, they increase the bulk hardness of
the cast iron simply by virtue of their own very high hardness and their substantial volume
fraction, such that the bulk hardness can be approximated by a rule of mixtures. In any case,
they offer hardness at the expense of toughness. Since carbide makes up a large fraction of the
material, white cast iron could reasonably be classified as a cermet. White iron is too brittle for
use in many structural components, but with good hardness and abrasion resistance and
relatively low cost, it finds use in such applications as the wear surfaces (impeller and volute)
of slurry pumps, shell liners and lifter bars in ball mills and autogenous grinding mills, balls and
rings incoal pulverisers, and the teeth of a backhoe's digging bucket (although cast mediumcarbon martensitic steel is more common for this application).
It is difficult to cool thick castings fast enough to solidify the melt as white cast iron all the way
through. However, rapid cooling can be used to solidify a shell of white cast iron, after which the
remainder cools more slowly to form a core of grey cast iron. The resulting casting, called
a chilled casting, has the benefits of a hard surface and a somewhat tougher interior.
High-chromium white iron alloys allow massive castings (for example, a 10-tonne impeller) to be
sand cast, i.e., a high cooling rate is not required, as well as providing impressive abrasion
resistance.[citation needed] These high-chromium alloys attribute their superior hardness to the presence
of chromium carbides. The main form of these carbides are the eutectic or primary
M7C3 carbides, where "M" represents iron or chromium and can vary depending on the alloy's
composition. The eutectic carbides form as bundles of hollow hexagonal rods and grow
perpendicular to the hexagonal basal plane. The hardness of these carbides are within the range
of 1500-1800HV[4]

Malleable cast iron[edit]

Main article: Malleable iron


Malleable iron starts as a white iron casting that is then heat treated at about 900 C (1,650 F).
Graphite separates out much more slowly in this case, so that surface tensionhas time to form it
into spheroidal particles rather than flakes. Due to their lower aspect ratio, spheroids are
relatively short and far from one another, and have a lower cross section vis-a-vis a propagating
crack or phonon. They also have blunt boundaries, as opposed to flakes, which alleviates the
stress concentration problems faced by grey cast iron. In general, the properties of malleable
cast iron are more like those of mild steel. There is a limit to how large a part can be cast in
malleable iron, since it is made from white cast iron.

Ductile cast iron[edit]


Main article: Ductile cast iron
A more recent development is nodular or ductile cast iron. Tiny amounts
of magnesium or cerium added to these alloys slow down the growth of graphite precipitates by
bonding to the edges of the graphite planes. Along with careful control of other elements and
timing, this allows the carbon to separate as spheroidal particles as the material solidifies. The
properties are similar to malleable iron, but parts can be cast with larger sections.
Comparative qualities of cast irons[5]

Name

Yield
Nominal
strength Tensile Elongatio Hardnes
compositio Form and
[ksi(0.2 strengt n [% (in s [Brinell
n [% by condition
%
h [ksi] 2 inches)] scale]
weight]
offset)]

Grey
C 3.4,
cast iron
Si 1.8, Mn 0. Cast
(ASTM A
5
48)

Uses

50

0.5

260

Engine cylinder
blocks, flywheels,gear
box cases, machinetool bases

C 3.4,
White
Si 0.7,
cast iron
Mn 0.6

Cast (as

cast)

25

450

Bearing surfaces

Malleabl
C 2.5,
e iron
Si 1.0,
(ASTM
Mn 0.55
A47)

Cast
(annealed 33
)

Ductile

C 3.4, P 0.1, Cast

53

52

12

130

Axle bearings, track


wheels,
automotivecrankshaft
s

70

18

170

Gears, camshafts,

or
Mn 0.4,Ni 1.
nodular
0, Mg 0.06
iron

Ductile
or
nodular

iron
(ASTM
A339)

Ni-hard
type 2

C 2.7,
Si 0.6,
Mn 0.5,
Ni 4.5,
Cr 2.0

C 3.0,
Si 2.0,
Ni-resist
Mn 1.0,
type 2
Ni 20.0,
Cr 2.5

crankshafts

cast
(quench 108
tempered)

135

310

Sand-cast

55

550

High strength
applications

Cast

27

140

Resistance to heat
and corrosion

How is cast iron classified?


Based on the alloying elements added, the variation in the solidification of the cast iron and heat treatment used,
the microstructure of the cast iron can vary. Depending upon the application and the preferred mechanical
properties, iron castings can be classified into the following.
Types of cast iron
White cast iron
When the white cast iron is fractured, white coloured cracks are seen throughout because of the presence of
carbide impurities. White cast iron is hard but brittle. It has lower silicon content and low melting point. The
carbon present in the white cast iron precipitates and forms large particles that increase the hardness of the cast
iron. It is abrasive resistant as well as cost-effective making them useful in various applications like lifter bars and
shell liners in grinding mills, wear surfaces of pumps, balls and rings of coal pulverisers, etc.
Grey cast iron
Grey is the most versatile and widely used cast iron. The presence of carbon leads to formation of graphite flakes
that does not allow cracks to pass through, when the material breaks. Instead, as the material breaks the

graphite initiates numerous new cracks. The fractured cast iron is greyish in colour, which also gives it the name.
The graphite flakes make the grey cast iron exhibit low shock resistance. They also lack elasticity and have low
tensile strength.
However, the graphite fakes gives the cast iron excellent machinability, damping features as well as good
lubricating properties making them useful in many industrial applications. The graphite microstructure of the cast
iron has a matrix that consists of ferrite, pearlite or a combination of two. The molten grey iron has greater fluidity
and they expand well during the solidification or freezing of cast iron. This has made them useful in industries like
agriculture, automobile, textile mills, etc.
Malleable cast iron
Malleable cast iron is basically white iron that undergoes heat treatment to convert the carbide into graphite. The
resultant cast iron has properties that vary from both grey and white cast iron. In case of malleable cast iron, the
graphite structure is formed into irregularly shaped spheroidal particles rather than flakes that are usually present
in gray cast iron. This make the malleable cast iron behave like low-carbon steel. There is considerable shrinkage
that results in reduced production of cast iron as well increased costs. Malleable cast iron can be identified easily
by the blunt boundaries.
Ductile cast iron
Ductile cast iron is yet another type of ferrous alloy that is used as an engineering material in many applications.
To produce ductile iron, small amount of magnesium is added to the molten iron, which alters the graphite
structure that is formed. The magnesium reacts with oxygen and sulphur in the molten iron leading to nodule
shaped graphite that has earned them the name-nodular cast iron. Like malleable iron, ductile iron is flexible and
exhibits a linear stress strain relation. It can be casted in varied sizes and into varying thickness.

Grey Iron vs. Nodular Iron


Many people will wonder the differences between grey iron and nodular iron, this article is intended to
introduce the grey iron and nodular iron in microstructure, physical properties and chemical
composition.

1. Microstructure
The key feature of grey iron is that it includes flakes of graphite which develop during the cooling
process. These graphite flakes give grey iron a distinctive gray color when it is fractured, and they are
also involved in many of the physical properties of this iron alloy.
The microstructure of nodular iron is the discrete form of the graphite nodules. Nodular iron is created
by an alloying process, which converts the crack-promoting graphite flakes of gray iron into nodules.
With this micro-structural transformation, the metal acquires superior ductility and elongation
characteristics.

2. Physical Property
Compared with grey iron, nodular iron has an absolute advantage in intensity. The max tensile
strength of nodular iron is 90k psi, while the max tensile strength of grey iron is only 35k psi.
Nodular Irons are generally superior to grey irons, regarding their yield strength. The max yield
strength of ductile iron is 40k psi; Grey iron is not very malleable or strong, it fractures easily.
Nodular iron is more flexible and elastic than other cast irons. Nodular iron has higher strengths,
greater elongation and better resistance to impact than grey iron.
The nodular iron family offers the design engineer a unique combination of strength, wear and fatigue
resistance and toughness as well as excellent ductility characteristics. In all its grades, nodular iron
exhibits mechanical properties that make it an ideal materials for mechanical and automotive parts.
Gray iron is the most versatile of all foundry metals. With the exception of wrought steel, grey iron is
the most widely used metal for engineering purposes. Grey iron is an extremely inexpensive metallic
material and is readily available in large quantities at almost any foundry.

3. Chemical Composition
Cast Iron Material

Chemical
composition

Grey iron (%)

Nodular iron (%)

2.9~3.5

3.5~3.8

Si

1.4~2.1

2.0~3.2

Mn

0.6~1.0

<0.5

0.1~0.5

<0.08

0.1~0.12

<0.025

Mg

0.03~0.06

RE
The different chemical composition of grey iron and nodular iron directly determines their different
properties. After reading the above analysis, we hope your can get a better understanding about grey
iron and nodular iron.

Austempered Ductile Iron


Abstract:

Austempered ductile iron (ADI) is finding an ever increasing worldwide market in the
automotive and other sectors. It offers a range of mechanical properties superior to
those of other cast irons, and shows excellent economic competitiveness with steels
and aluminum alloys.
ADI is a heat treated form of as-cast ductile iron. The heat treat process,
austempering, was developed with the intent of improving on the strength and
toughness of ferrous alloys. Ductile iron, with its relative low cost and ease to
manufacture, has been one of the largest beneficiaries of the austempering process.
Austempered ductile iron (ADI) is finding an ever increasing worldwide market in the automotive
and other sectors. It offers a range of mechanical properties superior to those of other cast irons,
and shows excellent economic competitiveness with steels and aluminum alloys.
ADI is a heat treated form of as-cast ductile iron. The heat treat process, austempering, was
developed with the intent of improving on the strength and toughness of ferrous alloys. Ductile
iron, with its relative low cost and ease to manufacture, has been one of the largest beneficiaries
of the austempering process. As a result, ADI has burst onto the scene in recent years with a
host of creative and innovative casting solutions.
The term "cast iron" designates an entire family of metals with a wide variety of properties. Cast
iron contains more than 2% carbon, present as a distinct graphite phase. In ductile cast iron the
graphite occurs as spheroids or spherulites rather than as individual flakes as in gray iron.
Ductile iron exhibits a linear stress-strain relation, a considerable range of yield strengths, and,
as its name implies, ductility.
"Austempering" is a high performance heat treatment for ferrous alloys which produces an
engineered, tailorable matrix structure. This austempered matrix structure gives tensile strength,
toughness, impact strength and fatigue properties that are comparable to heat-treated steels.
Figure 1 shows the tensile elongation of steel and austempered ductile iron ADI.

Figure 1: The tensile elongation between steels and austempered ductile iron

Austempered Ductile Iron (ADI)


Advantages
The ADI casting requires a precisely controlled heat-treatment (heat, old, quench, austemper,
and cool) to develop the desired microstructure (acicular ferrite and carbon-stabilized austenite)
and mechanical properties.
Properties of ADI compared to steel:

ADI is much easier to cast than steel


it is approximately 9% lighter than steel
it has minimal draft requirements compared with steel forgings
ADI loses less of its toughness than steel at sub-zero temperatures
work hardens when stressed
ADI has more damping capacity than steel.

The Austempering Process


Austempered ductile iron is produced by heat-treating cast ductile iron to which small amounts of
nickel, molybdenum, or copper have been added to improve hardenability. Specific properties are
determined by the careful choice of heat treating parameters. Austempering involves the
nucleation and growth of acicular ferrite within austenite, where carbon is rejected into the
austenite. The resulting microstructure of acicular ferrite in carbon-enriched austenite is called
ausferrite. Even though austenite in austempered ductile iron is thermodynamically stable, it can
undergo strain-induced transformation to martensite when locally stressed. The result is islands
of hard martensite that enhance wear properties.

Advanced Cast Products (ACP) uses salt baths for austenitizing, quenching, and austempering
in order to achieve close dimensional control. Times and temperatures are tightly controlled
throughout the entire process.
Steps in the ACP Austenitizing Process

1. Heat castings in a molten salt bath to austenitizing temperature.


2. Hold at austenitizing temperature to dissolve carbon in austenite.
3. Quench quickly to avoid pearlite.
4. Hold at austempering temperature in molten salt bath for isothermal transformation to
ausferrite.

1. Initial austenitizing times and temperatures (1550 to 1700F) are controlled to ensure
formation of fine grain austenite and uniform carbon content in the matrix. The precise
temperature is grade dependant.
2. Quench time must be controlled within a few seconds, to avoid the formation of pearlite around
the carbon nodules, which would reduce mechanical properties. Quench temperatures (450 to
750F) must stay above the point of martensite formation (except for ASTM A 897 Grade 5).
3. In the austempering step which follows austenitizing, the temperature of the final salt bath
must also be closely controlled. The austempering step is also precisely time-controlled, to avoid
over- or under-processing. By the end of this step, the desired ADI ausferrite structure has
developed.

Austempered ductile iron provides:

Yield strength, toughness and impact resistance comparable to many cast/forged steels.
Vibration dampening and heat transfer superior to other ferrous/non-ferrous alloys.
Significant weight and cost savings over both aluminum and steel castings/forgings.
Increased fracture and fatigue strength.
Cost savings over aluminum and steel castings/forgings.

Applications & Examples


The properties of ADI coupled with the cost and flexibility benefits of ductile iron castings means
the potential for ADI applications is vast:

Agriculture - excellent resistance to soil wear


Digger/Grab teeth - high strength and wear resistance
Industrial - wear components, pumps, etc.
Gears - for wear resistance and better vibration damping than steel
Construction - crushing, grading and wear components etc.
Food & feed milling - grinding, mixing, pelletizing etc.

Table 1: British Standards Specification for ADI EN 1564: 1997

Table 2: Standard ADI Grades (USA) ASTM 897-90 (ASTM 897M-90)

Date Published: Feb-2009

Introduction
What material offers the design engineer the best combination of low cost, design flexibility, good
machinability, high strength-to-weight ratio and good toughness, wear resistance and fatigue
strength? Austempered Ductile Iron (ADI) may be the answer to that question. ADI offers this superior
combination of properties because it can be cast like any other member of the Ductile Iron family, thus
offering all the production advantages of a conventional Ductile Iron casting. Subsequently it is
subjected to the austempering process to produce mechanical properties that are superior to
conventional ductile iron, cast and forged aluminum and many cast and forged steels.
Figures 4.1 and 4.2 compare the mechanical properties of ADI to those of conventional Ductile
Irons. Figure 4.1 also provides a comparison between the tensile strength-elongation relationships for
the ASTM A897 ADI specification and that of the ASTM A536 specification for conventional Ductile
Iron. These, and other Ductile Iron specifications are discussed in further detail in Section XII.

Compared to the conventional grades of Ductile Iron, ADI delivers twice the strength for a given level
of elongation. In addition, ADI offers exceptional wear resistance and fatigue strength.
Figure 4.3a shows the strength of Ductile Iron and ADI compared to cast and forged steels. Ductile
Iron has commercially replaced as cast and forged steels in the lower strength region, now ADI is
finding applications in the higher strength regions. As shown in Figure 4.3b, the yield strength of ADI is
over three times that of the best cast or forged aluminum. In addition ADI weighs only 2.4 times more
than aluminum and is 2.3 times stiffer. ADI is also 10% less dense than steel. Therefore, when you
compare the relative weight per unit of yield strength of ADI with that of various aluminums and steels
(Figure 4.4) it is easy to see the engineering and design advantages inherent in ADI.
For a typical component, ADI costs 20% less per unit weight than steel and half that of aluminum.
When we now analyze the cost-per-unit-strength of ADI vs. various materials (Figure 4.5) the
economic advantages of ADI become apparent.
The mechanical properties of Ductile Iron and ADI are primarily determined by the metal matrix. The
matrix in conventional Ductile Iron is a controlled mixture of pearlite and ferrite. (Tempered martensitic
matrices may be developed for wear resistance, but lack the ductility of either as-cast Ductile Iron or
ADI). The properties of ADI are due to its unique matrix of acicular ferrite and carbon stabilized
austenite; called Ausferrite. The austempering process is neither new or novel and has been utilized
since the 1930s on cast and wrought steels. The austempering process was first commercially
applied to Ductile Iron in 1972 and by 1998 world-wide production was approaching 100,000 tonnes
annually.
The preponderance of information on the austempering of steel and the superficial similarities
between the austempering heat treatments applied to steels and ADI, have resulted in comparisons
which are incorrect and damaging to the understanding of the structure and properties of ADI. ADI is
sometimes referred to as "bainitic Ductile Iron", but correctly heat treated ADI contains little or no
bainite. Bainite consists of a matrix of acicular (plate-like) ferrite and carbide. ADIs ausferrite matrix is
a mix of acicular ferrite and carbon stabilized austenite. This ausferrite may resemble bainite
metallographically, however it is not because it contains few or none of the fine carbides characteristic
in bainite. An ausferrite matrix will only convert to bainite if it is over tempered.
The presence of austenite in ADI also leads to a harmful misconception. It is "retained" in the sense
that it has persisted from the austenitizing treatment, but it is not the "retained austenite" that
designers and metallurgists equate with unstable, incorrectly heat treated steel. The austenite in ADI
has been stabilized with carbon during heat treatment and will not transform to brittle martensite even
at sub-zero temperatures.
The presence of stable, carbon enriched austenite also accounts for another inadequately understood
property of ADI. While thermodynamically stable, the enriched austenite can undergo a strain-induced
transformation when exposed to high, normal forces. This transformation, which gives ADI its
remarkable wear resistance, is more than mere "work hardening". In addition to a significant increase
in flow stress and hardness (typical in most metallic materials), this strain induced transformation also
produces a localized increase in volume and creates high compressive stresses in the "transformed"
areas. These compressive stresses inhibit crack formation and growth and produce significant
improvements in the fatigue properties of ADI when it is machined after heat treatment or subjected to
surface treatments such as shot peening, grinding or rolling. (See Section IX).
Back to Top
Mechanical Properties
ADI is a group of materials whose mechanical properties can be varied over a wide range by a
suitable choice of heat treatment. Figure 4.6 illustrates the strong correlation between austempering
temperature and tensile properties. A high austempering temperature, 750F (400C), produces ADI
with high ductility, a yield strength in the range of 500 MPa (72 ksi) with good fatigue and impact
strength. These grades of ADI also respond well to the surface strain transformation previously

discussed which greatly increases their bending fatigue strength. A lower transformation temperature,
500F (260C), results in ADI with very high yield strength (1400 MPa (200 MPa)), high hardness,
excellent wear resistance and contact fatigue strength. This high strength ADI has lower fatigue
strength as-austempered but it can be greatly improved with the proper rolling or grinding regimen.
Thus, through relatively simple control of the austempering conditions ADI can be given a range of
properties unequaled by any other material.
Tensile Properties
Like other Ductile Iron specifications presented in Chapter XII, ASTM A897 defines the minimum
tensile properties for different grades of ADI. Figure 4.1 indicates that the ranges of properties
exhibited by ADI exceed these minima, but does not offer quantitative evidence on which materials
selection decisions can be made with confidence. Competent producers of both conventional Ductile
Iron and ADI recognize that they must not only provide statistically significant mechanical property
data, but also give evidence of SPC and their commitment to continual improvement. Figure
4.7 provided by CMI International offers statistical evidence that grade 125/80/10 ADI can be
produced with mechanical properties significantly in excess of those required by the specification.
(The data shown represent 18 months of foundry production and over 600 heat treat lots).
The modulus of elasticity in tension for ADI lies in the range of 22.5-23.6 X 10 6 (155-163 GPa). Figure
4.8 shows the relationship of ADIs Youngs Modulus to that of other materials.
Fracture Toughness
Traditionally, components have been designed on the basis of preventing failure by plastic
deformation. As a result, design codes used either the 0.2% yield stress or the ultimate tensile stress
when specifying material properties, and designers then applied a safety factor when determining the
acceptable working stress level in the component. Both designers and metal casters have recognized
that structures also fail by brittle fracture and fatigue, especially in the presence of a crack-like defect.
As a result, fracture toughness, the intrinsic resistance of the material to crack propagation, is
becoming an essential part of the package of material properties used by designers to select
materials for critical applications.
There is a dearth of fracture toughness data for ADI, for two very valid reasons. First, being a
relatively new material, efforts to define a mechanical property database have concentrated on the
more conventional and easily acquired tensile properties, with early efforts at defining toughness
being confined largely to an extension of the notched and un-notched Charpy test used to
characterize conventional Ductile Irons. Second, the more ductile grades of ADI, for which fracture
toughness is a critical property, do not behave in a linear elastic manner when subjected to standard
LEFM tests, and toughness must be determined by yielding fracture mechanics techniques such as
the J integral crack opening displacement (COD) tests. Nevertheless, when the combined knowledge
of the toughness properties is assembled, they reveal three important facts:
1. Considering its high strength, ADI has very good toughness,
2. Where valid comparisons exist, the toughness of ADI is much greater than that of conventional
Ductile Iron and equivalent or superior to competitive cast and forged steels, and
3. Like other properties of ADI, its toughness is strongly dependent on microstructure (and thus, the
grade of ADI).
Figure 4.2 shows that ADI heat treated to produce high strength has a static fracture toughness of 5570 MPa(m)1/2, which is greater than that of Ductile Iron with a matrix of tempered martensite or
pearlite. Furthermore, it shows that ADI heat treated to a lower strength (higher ductility) grade has a
fracture toughness in excess of 100 MPa(m)1/2, twice as tough as pearlitic Ductile Iron. When
compared to forged steel and conventional Ductile Iron, both with equal or inferior mechanical
properties, ADI exhibits un-notched Charpy impact values at room temperature that are less than
those of forged steel, but three times higher than conventional Ductile Iron (Table 4.1). Figure

4.9 shows that the room temperature un-notched Charpy values of ADI are substantially higher than
those required by ASTM A897-90 at all levels of strength.
Table 4.1 Comparison of the mechanical properties of forged steel, pearlitic Ductile Iron and
Grade 150/100/7 ADI.
MATERIAL
Mechanical Property

Forged Steel Pearlitic Ductile Iron Grade 150/100/7 ADI

Yield strength, ksi (mPa)

75
(520)

70
(480)

120
(830)

Tensile strength, ksi (mPa)

115
(790)

100
(690)

160
(1100)

Elongation, %

10

10

Hardness, Bhn

262

262

286

130
(175)

40
(55)

120
(165)

Impact strength**, ft-lb (joules)

** Un-notched charpy at room temperature.


Table 4.2 provides a comparison of yield strength and fracture toughness, K1C, between ADI ,
conventional austenitic Ductile Irons, and quenched and tempered AISI 4140 and 4340 steels. With
K1C values in the range of 59-86 MPa m1/2, ADI had a fracture toughness which was superior to all
other Ductile Irons, except Ni-Resist, and equal to or higher than most of the quenched and tempered
steels. This table also compares the ratio of K1C to yield strength for these materials. This ratio, which
is proportional to the size of flaw that can be tolerated when materials are stressed to a constant
fraction of their yield strength, indicates that ADI has equal or greater flaw tolerance than pearlitic
Ductile Iron and quenched and tempered steels.
Table 4.2 Comparison of yield strength, fracture toughness and flaw tolerance between ADI,
conventional and austenitic Ductile Irons, and quenched and tempered steels.
Alloy

sy (MPa) KIC (MPa-m1/2) (KIC/sy)2 (mm)

Heat Treatment
o

A--2

850oC, 1 hr salt/quench

260 C*
300oC
350oC
400oC
430oC

1205.4
1107.4
989.8
744.8
744.6

73.49
68.62
72.10
72.91
74.52

3.72
3.84
5.30
9.58
10.00

B--5

850oC, 1 hr

260oC
300oC
350oC
400oC

1029.0
980.1
793.7
756.0

75.18
75.40
73.68
76.01

5.34
5.92
8.62
10.08

C--1

850oC, 1 hr

300oC

1151.5

86.00

5.58

C--3

850oC, 1 hr

300oC
350oC
400oC

1199.5
900.3
908.8

78.20
61.60
59.40

4.25
4.68
4.27

C--5

850oC, 1 hr

300oC

118.2

85.74

5.88

Ductile Iron, Ferritic, 1.55% Si, 1.5% Ni, 1.2% Ni

269

42.8

25.3

Ductile Iron, Ferritic, 3.6% C, 2.5% Si, 0.38% Ni, 0.35% Mo

331

48.3

21.3

Ductile Iron Pearlitic, 0.5% Mo

483

48.3

10.0

Ductile Iron, 80--60--03

432

27.1

3.9

Ductile Iron, D7003

717

51.7

5.2

Ductile Iron, Ni--Resist D-5B

324

64.1

39.1

204oC
280oC
396oC

1449
1587
1518

43.80
55.00
55.60

.92
1.2
1.34

1100oC, 1 hr., Oil Quench

204oC
246oC

1380
1449

65.05
57.25

2.22
1.56

1200oC, 1 hr., Oil Quench

204oC
246oC
323oC
348oC

1380
1449
1414.5
1393.8

89.12
72.64
53.30
58.46

4.18
2.52
1.42
1.76

(Above Data from: Iron Castings Handbook, 1981, p.357)


AISI
4140**

870oC, 1 hr., Oil Quench

200oC
280oC
350oC
400oC

1345
1504.2
1497.3
1449.0

65.38
66.81
87.69
100.22

2.36
1.97
3.43
4.78

1200oC, 1 hr., Oil Quench

246oC
280oC

1380.0
1393.8

90.55
69.01

4.31
2.45

843oC, 1 hr., OQ Tempered

260oC
427oC

1642.2
1421.4

48.79
84.47

.88
3.53

AISI 4340** 870oC, 1 hr., Oil Quench

* Isothermal Transformation Time: 1 hour.


** Data obtained from: Damage Tolerance Design Handbook, Metals and Ceramics Information, Battle
Columbus Laboratories, January 1975.
The ASME Gear Research Institute used both the ASTM Short Rod Fracture Test (SRFT) and the
non-standard Single Tooth Impact (STI) test to evaluate various ADI materials for gear applications
and compare their fracture toughness with carburized 8620 steel. Figure 4.10 shows ASTM Short
Rod Fracture Toughness for ADIs ranging from 55 to 105 MPa (m)1/2 at room temperature. This
compares favorably with a reported room temperature fracture toughness of 22 to 33 MPa (m) 1/2 for
carburized and hardened 8620 steel. The ADI toughness levels increase strongly with increasing
austempering temperature. (The austempering condition is indicated in degrees F and hours; i.e. ADI
750(.75 hr) was austempered at 750F for 45 minutes).
The Single Tooth Impact test results shown in Figure 4.11 are consistent with Short Rod Fracture
Toughness data, with ADI superior to carburized steel for both peak load and fracture energy criteria
at both room temperature and -40 degrees.
Figure 4.12 illustrates ADI dynamic fracture toughness data which were produced by instrumented
impact tests performed on Charpy-size specimens. The results, in the range of 40 to 80 MPa(m) 1/2,
show that toughness decreases with higher levels of manganese.
Regardless of the type of toughness test, ADI results were superior to those of conventional Ductile
Iron, and were equal to, or better than competitive steels. Additionally, all toughness tests revealed
that the toughness of ADI increases with austempering temperature to a maximum around 650-700F
(340-370C). Figure 4.13 confirms that this relationship is a further manifestation of the influence of
the volume fraction of stabilized austenite on the ductility and toughness of ADI.
An extensive work done at National University of Mar del Plata (Argentina) compared the properties of
ADI to those of 4140 steel. It concluded that while a standard notched Charpy test indicated that the
properties of ADI were inferior to those of 4140 steel, fracture toughness tests indicate "a much less
significant difference". In fact, it was found that the strain rate of the fracture test had a much more
significant effect on the steel than on the ADI. It concluded that "the comparison of toughness of ADI
and steels, should not be based on the impact energy measurements. Fracture mechanics properties,
such as K1C, should be used for design purposes".
Fatigue
As shown in Figure 4.14, ADI has fatigue properties equal or superior to those of forged steels. When
subjected to surface treatments such as rolling, peening or machining after heat treatment, the fatigue
strength of ADI is increased significantly. (See Figures 3.34, 3.35 and Table 3.3 in Section III,
and Figures 4.35 and 4.36. Figures 4.14 and 4.15 also indicate that ADI is moderately notch
sensitive in fatigue, with a notch sensitivity ratio (ratio of notched to un-notched endurance limits)
ranging from 1.2 to 1.6 (see Figure 4.16) for the notch geometry tested. Conventional ferritic and
pearlitic Ductile Irons have a notch sensitivity of about 1.6 and steels with fatigue strengths similar to
ADI exhibit notch sensitivity ratios as high as 2.2-2.4. To avoid problems caused by notch sensitivity,
components with sharp corners should be redesigned to provide generous fillets and radii. When
required, fillet rolling or shot peening can be employed to further increase resistance to fatigue failure.
Figure 4.15 relates the fatigue strengths of notched and un-notched ADI to tensile strength and
austempering temperature. Comparison of this figure with Figure 4.13 reveals several interesting

facts. Unlike conventional Ductile Iron, the un-notched fatigue limit of ADI does not follow the tensile
properties, demonstrates a maxima at the condition of lower tensile strength and maximum stabilized
austenite content in the metal matrix. These relationships result in an endurance ratio (ratio of fatigue
strength to tensile strength) that is 0.5 for lower strength ADI and decreases to 0.3 as the tensile
strength increases to its maximum. (See Figure 4.16). The notched ADI fatigue strengths shown
in Figure 4.15 increase rapidly with tensile strength and, as a result, high strength ADI is the least
notch sensitive.
Figure 4.17 shows an important relationship between austempering temperature and the endurance
limit of shot peened ADI. The dramatic rise in endurance limit in ADIs austempered above 600F
(315C) is related to the increased response to peening resulting from the higher austenite contents
characteristic of higher austempering temperatures.
Figures 4.18 and 4.19 indicate that for gear applications, shot peened ADI has single tooth bending
fatigue and contact fatigue superior to as cast and conventionally heat treated Ductile Irons, and cast
and through hardened steels. They also show that peened ADI is competitive with gas nitrided and
case carburized steels.
To this point we have discussed fatigue strength in terms that assume infinite life below a certain load.
In fact, as the number of loading cycles are increased all materials undergo changes in their ability to
withstand further loading. Figure 4.20 shows the relationship between number of cycles and the
allowable single tooth bending stress (in ksi). Grade 2 ADI represents an ADI austempered at 675F
(357C) and Grade 5 represents an ADI austempered at 500F (260C). Figure 4.21 shows the
relationship between the number of cycles and the allowable contact stress for those same
ADIs. Figure 4.22 shows the relationship of number of cycles to allowable rotating bending
stress for a Grade 1050 ADI (austempered at 675F (357C)) in the as-austempered condition.
To accurately model the finite element behavior of materials that are dynamically loaded the design
engineer uses certain coefficients and exponents to predict a components fatigue behavior. As of this
writing much work is being done to develop those numbers for Ductile Irons and ADI. Table 4.3shows
the typical fatigue coefficients and exponents for a 300 BHN ADI. (Courtesy of Ford Motor Company
and Meritor Heavy Vehicle Systems).
Table 4.3 Typical fatigue coefficients and exponents for 300 BHN ADI
Strength Coefficient K (ksi / MPa)
Strain Hardening Exponent n
True Fracture Strength Sf
True Fracture Ductility ef
Strength Coefficient K (ksi / MPa)
Strain Hardening Exponent n'
Fatigue Strength Coefficient sf' (ksi / MPa)

218 / 1503
0.143
150 / 1032
0.082
253 / 1744
0.1330
211 / 1455

Fatigue Strength Exponent b

-0.0900

Fatigue Ductility Coefficient ef'

0.1150

Fatigue Ductility Exponent C

-0.5940

Abrasion Resistance
Austempered Ductile Iron offers the design engineer abrasion resistance that is superior to
competitive materials over a wide range of hardness. Generally, ADI will outwear competitive
materials at a given hardness level. For example (from Figure 4.23) an ADI component at 30 to 40 Rc
will wear comparably to a quenched and tempered steel component at nearly 60 Rc in an abrasive
wear environment. This property, shown in Figures 4.23 and4.24, allows the designer to select the
combination of strength, ductility and abrasion resistance that will provide the best component
performance in a particular application.

The superior abrasion resistance, and the low sensitivity of abrasion resistance to bulk hardness are
related to the strain-induced transformation of stabilized austenite which occurs when the surface of
an ADI component is subjected to deformation. The result of this transformation is a significant
increase in surface hardness shown in Figure 4.25. This increase in surface hardness, and its
relationship to microstructure, are responsible for the reduced sensitivity of abrasion resistance to
hardness. As the bulk hardness of ADI is reduced by the austempering temperature, the amount of
stabilized austenite increases (see Figure 4.26). This increase in austenite content increases the
hardness increment produced by surface deformation. As a result, a Ductile Iron component
austempered to produce a lower hardness displays an abrasion resistance greater than that predicted
by its bulk hardness, provided that the abrasion mechanism involves sufficient deformation to
transform the surface layers to martensite.
Through variations in austempering conditions, the designer can optimize the abrasion resistance and
related mechanical properties of an ADI component. For a combination of high toughness and
abrasion resistance an austempering temperature in the range of 650-700F (350-375C) should be
used. When a combination of high strength and abrasion resistance are required, an austempering
temperature of 500F (260C) will yield the best results.
Machinability
Compared to competitive materials, the machinability of Ductile Iron has been one of its major
advantages. When the substantial increases in strength and wear resistance offered by ADI are
considered, it would be logical to assume that ADI could present machining problems. However, cost
savings in machining are frequently mentioned as reasons for converting to ADI. The reasons for this
surprising combination of mechanical properties and machinability are two-fold. First, the
machinability of the softer grades of ADI is equal or superior to that of steels with equivalent strength,
and second, the predictable growth characteristics of ADI during austempering allow, in many cases,
for it to be machined complete in the soft as-cast or annealed state before heat treatment. This allows
for faster machine feeds and speeds and greatly increased tool life. (See Figure 4.27).
As discussed in a later section on surface deformation treatments of gears, the processing of ADI
parts should could follow one of several paths, (Figure 4.34), depending primarily on the grade of
ADI, but also on the surface treatment benefits that may be gained from machining after
austempering. As shown in Figures 4.28 and 4.29, cutting tool life decreases substantially as the
hardness of ADI increases and the cutting speed and metal removal rate increase. For these reasons,
only the 125/80/10 and 150/100/07 grades of ADI should be machined after austempering. Parts
processed to the higher strength grades should receive the following processing sequence:
1. Cast the component
2. (Optionally) subcritically anneal to a fully ferritic matrix
3. Machine
4. Austemper
5. Finish machine (if required)
6. Finish operations (rolling, grinding, peening, if required)
While annealing adds cost to the casting the benefits of more predictable dimensional change during
austempering and greatly improved machinability often more than offset the cost of annealing. In
order to obtain the benefits of the excellent machinability of annealed Ductile Iron, the designer must
have confidence in the reproducibility of the growth of the machined casting during heat treatment.
Papers presented at the 1st International Conference on ADI indicated that dimensional changes
during heat treatment varied from a slight contraction to a growth of approximately 0.4%, (Figure
4.30). It was stressed that as long as the prior Ductile Iron microstructure was consistent in
pearlite/ferrite ratio, predictable growth occurred and close tolerance ADI parts could be produced
successfully by machining prior to heat treatment.

The production of ADI ring and pinion gear sets for General Motors rear wheel drive cars (model years
1977-1979) provides a good example of the reduction in machining costs offered by ADI. As shown
in Table 4.4, tool life improvement for different machining operations performed on the ferritized ADI
blanks ranged from 20% to over 900% compared to similar operations required to produced the
forged, carburized and hardened 8620 steel gears that they replaced. The overall cost savings to GM
of converting from carburized steel to austempered ring and pinion gears was approximately 20%.
Table 4.4 Tool life improvement resulting from the replacement of casecarburized forged steel
gears with ADI.
Machining operation Tool-life improvement %
Pinion blanking
- centre press
- drill
- rough lathes
- finish lathes
- grind

30
35
70
50
20

Rear-gear blanking
- bullard turning
- drilling
- reaming

200
20
20

Gleason machining
- pinion - roughing
- pinion - finishing
- ring - roughing
- ring - finishing

900
233
962
100

Tables 4.5 and 4.6 provide guidelines for the machining of the lower strength grades of ADI. In
general, reduced surface speeds and increased feed rates provide the best metal removal rate in ADI.
Table 4.5 Guidelines for the machining of a lower strength grade of ADI.
Machining operation Tools

Cutting speed, m/min Feed, mm/revolution

Turning

K20 bits with TiC


angle g=-6o,
no cutting-oil,
tool force 1.6-1.8 kN/mm2

50 - 70

Roughing: 0.5 - 1.0


Finishing: 0.15 - 0.3

Drilling

Carbide-tipped drills

12 - 15

0.05 - 0.3

Keyway broaching

High-speed steel,
g = 10o, a = 5o

3-6

0.05 - 0.08

8 - 20

1.5 - 2.5
depending on module

with cutting oil

Hobbing

Hobs flooded with cutting-oil

Grinding

Grade 37C16-P4B wheels

Table 4.6 Recommended machining conditions for a lower strength grade of low-manganese
ADI.
Turning
Machining
Cutting-speed, m/min
Feed, mm
Depth of cut, mm
Wear, mm
Tool life, min
Lubrication

Drilling

High-speed steel SiN2 High-speed steel Carbide


100
0.355
2
0.7
12
Yes

100
.25
4
0
>10
Yes

20
.25
12.5 dia
20
Yes

50
.18
11.5 dia
25
Yes

Production Control
The mechanical properties offered by ADI make it an attractive material for demanding applications in
which strict specifications must be met consistently. While greatly enhancing the properties of a
conventional Ductile Iron casting, the ADI process cannot compensate for casting defects that would
impair mechanical properties. ADI castings should, therefore, be produced free from surface defects,
and with the following microstructural parameters.

Nodularity: >80% type I and II nodules


Nodule Count: 100/mm2 minimum

Consistent chemical composition

Essentially free of carbides, porosity and inclusions

Consistent pearlite/ferrite ratio

These requirements are essentially the same as those required to produce good quality Ductile Iron.
To assure quality, ADI should be purchased from casting and heat treatment suppliers that have well
developed process control systems who can demonstrate that they are consistently capable of
producing high quality castings and heat treatments.
The production of a high quality casting is essential but, by itself, not a sufficient condition to ensure
optimum properties in ADI. The casting must be heat treated properly by a supplier capable of taking
into account the interaction between casting dimensions, composition, microstructure and the desired
properties in the austempered casting. For this reason, there should be close cooperation between
the designer, metal caster, heat treater and machine source from conception of the design to delivery
of the castings.
Composition
In many cases, the composition of an ADI casting differs little from that of a conventional Ductile Iron
casting. When selecting the composition, and hence the raw materials, for both conventional Ductile
Iron and ADI, consideration should be given first to limiting elements which adversely affect casting
quality through the production of non-spheroidal graphite, or the formation of carbides and inclusions,
or the promotion of shrinkage. The second consideration is the control of carbon, silicon and the major
alloying elements (See Table 4.7) that control the hardenability of the iron and the properties of the
transformed microstructure. When determining the alloying requirements both the section size and
type and the severity (or speed) of the austempering quench must be considered.
For a typical salt quench with agitation section sizes up to about 3/8 inch (10 mm) can be successfully
through hardened without pearlite with even unalloyed Ductile Iron. For a highly agitated austemper
quench with water saturation section sizes of up to inch (20 mm) can be through hardened with no
additional alloying. For castings of heavier section size selective alloying is required to through harden
the parts and avoid pearlite in the heat treated microstructure.
Figure 4.31 summarizes the hardenability of copper, nickel and molybdenum by relating the levels of
these elements to the maximum diameter bar that can be satisfactorily austempered. (The BCIRA
data utilizes a relatively fast quench while the Dorazil data reflects an austempering bath with a lower
quench severity). The relative hardenability contribution of manganese is between that of
molybdenum and copper.
Table 4.7 Effects of carbon, silicon and the major alloying elements an austempering behavior.

Carbon

Increasing carbon in the range 3 to 4% increases the tensile strength but has negligible
effect on elongation and hardness. Carbon should be controlled within the range 3.63.8% except when deviations are required to provide a defect-free casting.

Silicon

Silicon is one of the most important elements in ADI because it promotes graphite
formation, decreases the solubility of carbon in austenite, increases the eutectoid
temperature, and inhibits the formation of bainitic carbide. Increasing the silicon content
increases the impact strength of ADI and lowers the ductile-brittle transition temperature.
Silicon should be controlled closely within the range 2.4-2.8%.

Manganese can be both a beneficial and a harmful element. It strongly increases


hardenability, but during solidification it segregates to cell boundaries where it forms
carbides and retards the austempering reaction. As a result, for castings with either low
nodule counts or section sizes greater than 3.4 in. (19mm), manganese segregation at
cell boundaries can be sufficiently high to produce shrinkage, carbides and unstable
Manganese austenite. These microstructural defects and inhomogeneities decrease machinability
and reduce mechanical properties. To improve properties and reduce the sensitivity of
the ADI to section size and nodule count, it is advisable to restrict the manganese level
in ADI to less than 0.3%. The use of high purity pig iron in the ADI charge offers the twin
advantages of diluting the manganese in the steel scrap to desirable levels and
controlling undesirable trace elements.
Copper

Up to 0.8% copper may be added to ADI to increase hardenability. Copper has no


significant effect on tensile properties but increases ductility at austempering
temperatures below 675oF (350oC).

Nickel

Up to 2% nickel may be used to increase the hardenability of ADI. For austempering


temperatures below 675oF (350oC) nickel reduces tensile strength slightly but increases
ductility and fracture toughness.

Molybdenum is the most potent hardenability agent in ADI, and may be required in heavy
section castings to prevent the formation of pearlite. However, both tensile strength and
ductility decrease as the molybdenum content is increased beyond that required for
Molybdenum
hardenability. This deterioration in properties is probably caused by the segregation of
molybdenum to cell boundaries and the formation of carbides. The level of molybdenum
should be restricted to not more than 0.2% in heavy section castings.
Alloy combinations
For economic reasons, or to avoid metallurgical problems, combinations of alloys are often used to
achieve the desired hardenability in ADI. To avoid micro-segregation and the resultant degradation of
mechanical properties associated with higher levels of manganese and molybdenum, their levels
should be carefully controlled with the desired hardenability obtained by supplementary additions of
first copper (up to about 0.8%), then nickel.
The Role of Alloys in ADI
The primary purpose of adding copper, nickel or molybdenum to ADI is to increase the hardenability of
the matrix sufficiently to ensure that the formation of pearlite is avoided during the austempering
process. These elements have only marginal effect on the mechanical properties of ADI that is
properly austempered. (The austempering process determines the properties after austempering; not
the alloying). Only the minimum amount of alloys required to through harden the part should be
employed. Excessive alloying only increases the cost and difficulty of producing the good quality
Ductile Iron necessary for ADI. Ultimately, the amount of alloying required will be a function of the
metal casters base composition, the casting section size and type and the characteristics of the
chosen heat treatment process.
A typical iron composition (and control range) that can be used is shown below:
Carbon*

3.7%

+/- 0.2%

Silicon*

2.5%

+/- 0.2%

Manganese

0.28%

+/- 0.03%

Copper

as required

+/- 0.05% up to 0.8%


maximum

Nickel

as required

+/- 0.10% up to 2.0%


maximum

Molybdenum

only if required

+/- 0.03% up to 0.25%


maximum

*(Carbon and silicon should be controlled to produce the desired carbon equivalent for the section
size being produced).
The aforementioned composition does not guarantee ADI properties, nor is it mandatory. However,
this composition is a typical, industrially successful ADI composition. A good controlled chemistry like
this one combined with a consistently high nodule count and nodularity and a consistent
pearlite/ferrite ratio in clean, shrink-free Ductile Iron will provide the most robust process for the
production of ADI.
Heat Treatment
ADI is produced by an isothermal heat treatment known as austempering. Austempering consists of
the following steps as shown in Figure 4.32:
1. Heating the casting to the austenitizing temperature in the range of 1500-1700F (815-927C).
2. Holding the part at the austenitizing temperature for a time sufficient to get the entire part to
temperature and to saturate the austenite with carbon.
3. Quenching (cooling) the part rapidly enough to avoid the formation of pearlite to the
austempering temperature in the range of 450-750F (232-400C). (This temperature is above
the martensite start temperature (Ms) for the material)
4. Austempering the part at the desired temperature for a time sufficient to produce a matrix of
ausferrite. (That is a matrix of acicular ferrite and austenite stabilized with about 2% carbon).
5. Cooling the part to room temperature
The austenitizing may be accomplished by using a high temperature salt bath, an atmosphere furnace
or (in special cases) a localized method such as flame or induction heating. The austempering is most
typically carried out in a nitrite/nitrate salt bath but in special cases it can be accomplished in hot oil
(up to 470F (243C)), or molten lead or tin.
The critical characteristics are:

The austenitizing time and temperature


A cooling rate sufficient for the casting/alloy combination

The austempering time and temperature

Austenitizing
The austenitizing temperature controls the carbon content of the austenite which, in turn, affects the
structure and properties of the austempered casting. High austenitizing temperatures increase the
carbon content of the austenite, increasing its hardenability, but making transformation during
austempering more problematic and potentially reducing mechanical properties after austempering.
(The higher carbon austenite requires a longer time to transform to ausferrite). Reduced austempering
temperatures generally produce ADI with the best properties but this requires close control of the
silicon content, which has a significant effect on the upper critical temperature of the Ductile Iron.

Austenitizing time should be the minimum required to heat the entire part to the desired austenitizing
temperature and to saturate the austenite with the equilibrium level of carbon, (typically about 1.11.3%). In addition to the casting section size and type, the austenitizing time is affected by the
chemical composition, the austenitizing temperature and the nodule count.
Austempering
Austempering is fully effective only when the cooling rate of the quenching apparatus is sufficient for
the section size and hardenability of the component. The minimum rate of cooling is that required
avoid the formation of pearlite in the part during quenching to the austempering temperature. The
critical characteristics are as follows:

Transfer time from the austenitizing environment to the austempering


environment
The quench severity of the austempering bath

The maximum section size and type of casting being quenched

The hardenability of the castings

The mass of the load relative to the quench bath.

The use of a correctly designed austempering system with a suitably high quench severity, and the
correct loading of castings, can minimize hardenability requirements of the casting resulting in
significant savings in alloy costs.
As illustrated earlier in Figure 4.6, austempering temperature is one of the major determinants of the
mechanical properties of ADI castings. To produce ADI with lower strength and hardness but higher
elongation and fracture toughness, a higher austempering temperature (650-750F (350-400C)) should
be selected to produce a coarse ausferrite matrix with higher amounts of carbon stabilized austenite
(20-40%). Grades 125/80/10 and 150/100/07 would be typical of these conditions. To produce ADI
with higher strength and greater wear resistance, but lower fracture toughness, austempering
temperatures below 650F (350C) should be used.
Once the austempering temperature has been selected, the austempering time must be chosen to
optimize properties through the formation of a stable structure of ausferrite. Figure 4.33 schematically
illustrates the influence of austempering time on the stabilization of austenite, and shows the
hardness of the resultant matrix. At short austempering times, there is insufficient diffusion of carbon
to the austenite to stabilize it, and martensite may form during cooling to room temperature. The
resultant microstructure would have a higher hardness but lower ductility and fracture toughness
(especially at low temperatures). Excessive austempering times can result in the decomposition of
ausferrite into ferrite and carbide (bainite) which will exhibit lower strength, ductility and fracture
toughness. At the highest austempering temperature (750F (400C)) as little as 30 minutes may be
required to produce ausferrite. At 450F (230C) as much as four hours may be required to produce the
optimum properties; Figure 4.32.
Note: a strength level maxima is achieved in ADI at an austempering temperature of about 475-525F
(250-275C). At temperatures below that range the hardness may increase but the strength may
decrease due to the presence of martensite mixed in with the ausferritic matrix. (In other words, as the
austempering temperature is incrementally decreased below 475F (250C) the material behaves
increasingly like a quenched and tempered Ductile Iron).
Surface Deformation Treatments
The austempering treatment used to produce ADI can result in small residual tensile surface stresses.
Even so, ADI has excellent fracture toughness and fatigue strength. However, ADIs ability to resist
the initiation of fatigue cracks (especially in cyclic bending) may be greatly enhanced by inducing
compressive stresses to the surface after the austemper heat treatment.

These small tensile stresses can be easily replaced with rather substantial compressive stresses if the
ADI is subjected to any surface treatment involving the sufficient surface deformation to cause a
strain-induced transformation of the stabilized austenite. Such treatments could include conventional
machining operations such as turning, grinding, milling, hobbing or special treatments such as shot
peening or surface rolling.
For parts subjected to fatigue failure, performance can be enhanced significantly if machining
operations can be performed after austempering. However, this processing sequence is limited by the
machinability of the different grades of ADI. Figure 4.34 shows the two processing flow charts used
by the ASME Gear Research institute to produce ADI test gears. Using machinability as the main
criterion, the GRI selected process #2 for blanks with heat treated hardnesses in the range of 30-34
HRC and process #1 for blanks with hardness greater than 37 HRC. Figure 4.35 reveals that hobbing
after austempering resulted in a 20% increase in fatigue strength of gear teeth.
Shot Peening
Shot peening offers a controllable means of selectively hardening certain parts of a finished casting to
produce significant improvements in fatigue properties. Figure 4.35 shows that shot peening a
hobbed gear increased fatigue strength by 60%. Single tooth fatigue tests conducted by the GRI on
ADI gears indicated that shot peening doubled the fatigue strength. The GRI work also identified a
significant correlation between residual compressive stress produced by peening and the endurance
limit (Figure 4.36). In addition to being able to be applied to selected areas of a part, peening also
offers the advantage of increasing surface hardness without detrimentally affecting the ductility and
strength of the remainder of the component. Post peening surface treatments such as honing may be
required to reduce surface roughness in parts subjected to rolling contact fatigue. (For more
information seeSection IX).
Surface Rolling
Surface rolling or burnishing produced with a hardened roller, or mating part in the case of gear
hardening, may be used independently or as a post-peening operation to improve the fatigue
properties of ADI. Gear Research Institute tests show that roll burnishing of a peened gear with a
carburized steel mate produced a six-fold increase in fatigue life. Fillet rolling has been used very
effectively to offset reductions in fatigue life produced by stress concentrations at changes in cross
section in castings such as crankshafts and gear clusters (see Table 3.3 and Section IX).
Other Properties
There are many properties needed for specific design applications that are not currently published.
Without these important properties engineers must make assumptions about a materials behavior
and mistakes can be made, (as discussed in some of the referenced papers listed in this chapter).
Table 4.9 summarizes some of those properties for the five ASTM897 grades.
Table 4.9
ASTM 897 (in-lb units)

125 80 10

150 100 07

175 125 04

200 155 01

230 185 00

ASTM 897M (SI units)

850 550 10

1050 700 07

1200 850 04

1400 1100
01

1600 1300 00

Min. Tensile Strength (ksi/MPa)

125/850

150/1050

175/1200

200/1400

230/1600

Min. 0.2% Offset Yield Strength (ksi/MPa)

80/550

100/700

125/850

155/1100

185/1300

10

n/a

302 (3.50)

340 (3.30)

387 (3.10)

418 (3.00)

460 (2.85)

"Grade"
Property

Min. Elongation (% in 2 in/50mm gage)


Typical Brinell Hardness (BID mm)
3

Typical Density (lb./in / g/ cm ) .2562/7.0965 .2555/7.0779 .2555/7.0779 .2552/7.0686 .2548/7.0593

Typical Thermal expansion (in/in/F /


mm/mm/C)

8.1/14.6

Typical Thermal conductivity


153/22.1
(BTU-in/h-ft2 / W-MK)
Typical Internal Damping (log
decrement X .0001)

5.26

8.0/14.3

7.8/14.0

7.7/13.8

7.5/13.5

151/21.8

149/21.5

147/21.2

145/20.9

5.41

5.69

12.7

19.2

Applications
The development and commercialization of Austempered Ductile Iron (ADI) has provided the design
engineer with a new group of cast ferrous materials which offer the exceptional combination of
mechanical properties equivalent to cast and forged steels and production costs similar to those of
conventional Ductile Iron. In addition to this attractive performance: cost ratio, ADI also provides the
designer with a wide range of properties, all produced by varying the heat treatment of the same
castings, ranging from 10-15% elongation with 125 ksi (870 MPa) tensile strength, to 250 ksi (1750
MPa) tensile strength with 1-3% elongation. Although initially hindered by lack of information on
properties and successful applications, ADI has become an established alternative in many
applications that were previously the exclusive domain of steel castings, forgings, weldments,
powdered metals and aluminum forgings and castings.
The ADI market represents nearly all segments of manufacturing. Figure 4.37 shows the approximate
breakdown of the North American ADI market.
Heavy Truck and Bus Components
Heavy truck applications include suspension components such as spring hanger brackets, shock
brackets, u-bolt plates, wheel hubs, brake calipers and spiders, knuckles, sway bar components,
pintle hooks and gears for trailer landing gear. Powertrain related ADI heavy truck and bus
components include engine brackets and mounts, timing gears, cams, annular gears, differential
gears and cases, clutch collars, accessory brackets and pulleys.
Light Auto and Truck Components
Light vehicle ADI applications include suspension components, knuckles, spindles, hubs, tow hooks,
hitch components, differential gears and cases, engine and accessory brackets, camshafts, engine
mounts, crankshafts and control arms. Constant velocity joints for four wheel drive GM vehicles have
been produced in ADI since 1978 and currently run at volumes of over 5,000 per day.

Truck spring support in ADI which replaced cast steel (Muhlberger GGG100B/A)
Construction and Mining Components
This segment includes all manner of collets, ring carriers, wear plates, sprockets, covers, arms,
knuckles, shafts, rollers, track components, tool holders, digger teeth, cutters, mill hammers, cams,
sway bars, sleeves, pavement breaker bodies and heads, clevises, and conveyor components.

Assorted ADI conveyor components


Miscellaneous Industrial
Miscellaneous industrial applications include brackets, lever arms, knuckles, shafts, cams, sway bars,
sleeves, clevises, conveyor components, jack components, bushings, rollers, molding line
components, fixtures, gears, sprockets, deck plates, and all sorts of power transmission and structural
components
Railroad
The railroad industry uses ADI for suspension housings, top caps and friction wedges, track plates,
repair vehicle wheels, nipper hooks, and car wheels.

Agricultural
Farming and agricultural applications for ADI include plow points, till points, trash cutters, seed boots,
ammonia knives, gears, sprockets, knotter gears, ripper points, tractor wheel hubs, rasp bars, disk
parts, bell cranks, lifting arms, and a great variety of parts for planters, plows, sprayers and
harvesters.
Defense
The defense industry has been relatively slow to adopt ADI, however some of the applications include
track links, armor, ordnance and various hardware for trucks and armored vehicles.
Special Product Categories

Gears
Gears represent some of the best known, most widely publicized and high potential uses of ADI.
During the early 1970s the Finnish company Kymi Kymmene Metall began to replace forged steel
with ADI in a in a wide range of gears, with highly satisfactory results.
In North America, ADI achieved a major breakthrough in 1977, when General Motors converted a
forged and case hardened steel ring gear and pinion to ADI for Pontiac rear drive cars and station
wagons. The decision came after nine years of development work and six years of field testing. The
automaker was able to gain both significant cost savings and product improvement by changing to
ADI.

ADI Timing Gears for Cummins B-Series diesel engines. Replaced forged and case
carburised 1022 steel with 30% cost saving.
In 1983, the Cummins Engine Co. began to use ADI timing gears, produced to AGMA class 8
standards, in its B and C series diesel engines. These gears were machined and hobbed from
annealed Ductile Iron castings. A crown shaving operation was carried out on the gear teeth prior to
austempering, and the only operations performed after austempering were the grinding of the bore
diameter and shot peening. Annual production exceeds 30,000 sets and the cost savings are
estimated at 30% compared to the forged and carburized 1022 steel gears previously used.
Table 4.8 Comparison of energy requirements for the production of ADI and forged and
carburized steel gears.
Energy consumption
kWh per Tonne
Operation

ADI

Production of blank 2,500


Annealing

Forged steel
4,500
500

Austempering

600

---

Case-hardening

---

800-1,200

Total

3,100

5,800-6,200

Table 4.8 describes the energy savings of almost 50% resulting from the conversion to ADI gears. In
addition to savings in energy and overall production costs, ADI gears offer the following advantages:

increased machine shop productivity


reduction in weight of up to 10%

reduced gear noise

rapid "break in" of new gears and,

improved resistance to scoring.

Austempered Ductile Iron gears to patented specifications K9805.


Courtesy: Hymi Kymmene Engineering, Finland.

Austempered Ductile Iron Hypoid Axle Gears: Conversion to Cast Ductile Iron from Forged
Steel gave:
major production cost saving, better machinability, quieter operation, reduced weight.
Courtesy: General Motors Corp. Central Foundry Div., Saginaw, Michigan, USA
Crankshafts
Crankshafts are another potentially significant application for Austempered Ductile Iron. The first
commercially produced ADI part was a small crankshaft for a hermetically sealed refrigerator
compressor. It was cast by Wagner Castings Company (US) for Tecumseh Products. Production of
that part was initiated in 1972 and since that time, millions of those crankshafts have been produced.

Engines being developed by the automotive industry require weight reduction in parts that will be
required to handle increased power. Automotive design engineers have evaluated ADI as a candidate
for both the replacement of forged steel crankshafts and the upgrading of existing Ductile Iron
crankshafts. The Ford Motor Company made an exhaustive, three year study of ADI crankshafts and
concluded that they met all design criteria. During this study, the importance of fatigue testing was
identified, and the following results were obtained:
Fatigue Test Method

Fatigue Strength
ksi

MPa

Constant Strain Amplitude

55

380

Rotating Bending

65

450

Reversed Bending

60

415

Reversed Bending (fillet rolled)

90

620

A thorough, joint Motor Industry Research Association / Cast Metals Development Laboratories study
on ADI crankshafts concluded that properly fillet rolled ADI crankshafts exhibited fatigue properties
comparable to, or better than, the best forged and heat treated steel crankshafts.
In another documented crankshaft study conducted at the Manchester (England) Materials Science
Center, the authors demonstrated the performance capability of ADI crankshafts in one cylinder
commercial and four cylinder automotive engines. They noted a 10% rotating weight reduction and an
estimated 30% cost savings.
As of this writing ADI crankshafts are employed in high volume commercial applications and low
volume automotive applications. As the specific power requirements for automotive engines are
increased, ADI will become a more viable alternative to the heavier, more expensive forged steel
crankshaft.

An assortment of ADI "ground engaging" parts. Courtesy of Applied Process Inc.

ADI crawler type track shoe. Courtesy of Applied Process Inc.

ADI crankshaft for a hermetically sealed compressor (first produced in 1972).


Courtesy of Wagner Castings and Tecumsen Products.
Conclusion
As design engineers become more familiar with ADIs strength, toughness, wear resistance and noise
damping properties and learn about the impressive cost and weight savings reported in successful
ADI conversions from steel castings, weldments and forgings and aluminum castings and forgings,
ADI will continue its remarkable growth.

Cast Irons

Cast irons like steels are basically alloys of iron and carbon. Cast irons contain between 2 and 6.67
percent carbon. Since the high carbon content tends to make cast irons very brittle, most
commercially manufactured types are in the range of 2.5 to 4.0 percent carbon.

Generally the ductility of cast iron is very low, and it cannot be rolled, drawn or worked at room
temperature. However, they melt readily and can be cast into complicated shapes which are
usually machined to final dimensions. They are known as cast irons because casting is the only
suitable process applied to these alloys. Though the common cast irons are brittle and have lower
strength, they can be cast more readily than steels and are cheap.
Cast iron is brittle. is an outdated but widely held truism which mistakenly implies that all cast
irons are the same, and none are ductile. In fact, they can be made ductile. Malleable cast iron
made from white cast iron and nodular iron are ductile. Ductile iron offers the design engineer a
unique combination of a wide range of high strength, wear resistance, fatigue resistance,
toughness and ductility in addition to the well-known advantages of cast iron castability,
machinability, damping properties, and economy of production. Unfortunately, these positive
attributes of ductile iron are not as widely known as the mistaken impression of brittleness is well
known.
Information about various types of cast iron is given in this article.

Classification of Cast Irons


The best method of classifying cast iron is according to metallographic structure.
There are four variables as under to be considered which lead to the different types of cast irons.
1.

Carbon content

2.

The alloy and impurity content

3.

The cooling rate during and after freezing

4.

The heat treatment after casting.

Above variables control the condition of the carbon and also its physical form. The carbon may be
combined as iron carbide in cementite, or it may exist as free carbon in graphite. The shape and
distribution of the free carbon particles will greatly influence the physical properties of the cast
iron. Various types of cast irons are as under.

White cast irons


In this type of cast irons, all the carbon is in the combined form as cementite.

Malleable cast irons


In this type of cast irons, most or all of the carbon is uncombined in the form of irregular round
particles known as temper carbon. This is obtained by heat treatment of white cast iron.

Gray cast irons


In this type of cast irons, most or all of the carbon is uncombined in the form of graphite flakes.

Chilled cast irons


In this type of cast irons, a white cast iron layer at the surface is combined with a grey cast iron
interior.

Nodular cast irons


By special alloy additions, in this type of cast irons, the carbon is largely uncombined in the form of
compact spheroids. This structure differs from malleable iron in that it is obtained directly from
solidification and the round carbon particles are more regular in shape.

Alloy cast irons


In this type of cast irons, the properties or the structure of any of the above types are modified by
the addition of alloying elements.
Detail information on above cast irons is given below.

White Cast Irons

All white cast irons are hypoeutectic alloys. Their typical microstructure on fast cooling from the
liquid state consists of dendrites of transformed austenite (pearlite) in a white interdendritic
network of cementite. Since white cast iron contains a relatively large amount of cementite as a
continuous interdendritic network, it makes it hard and wear resistant but extremely brittle and
difficult to machine. Completely white cast irons are limited in engineering applications due to this
brittleness and lack of machinability. They are used where resistance to wear is most important and
the service does not require ductility, such as liners for cement mixer and extrusion nozzles. A
large quantity of white cast iron is used as a starting material for the manufacture of malleable
cast iron. The range of mechanical properties for unalloyed white cast iron is as under.
Tensile strength: 20000 to 70000 psi
Hardness, Brinell: 375 to 600

Malleable Cast Irons

Cementite (iron carbide) is a metastable phase. There is a tendency for cementite to decompose
into iron and carbon, but under normal conditions it tends to persist indefinitely in its original form.
This tendency to form free carbon is the basis for the manufacture of malleable cast iron.
Decomposition of cementite into iron and carbon is favored by elevated temperature, the existence
of solid nonmetallic impurities, higher carbon contents and the presence of elements that aid the
decomposition of cementite.

The purpose of malleabilization is to convert all the combined carbon in white iron into irregular
nodules of temper carbon (graphite) and ferrite. Commercially, this process is carried out in two
steps known as the first and second stages of the anneal.
White irons suitable for conversion to malleable iron are of the following range of composition.
Carbon: 2.00-2.65 percent
Silicon: 0.90-1.40 percent
Manganese: 0.25-0.55 percent
Phosphorus: Less than 0.18 percent
Sulfur: 0.05 percent
In the first stage of annealing, the white iron casting is slowly reheated to a temperature between
1650 and 1750F. They are held at this temperature until all massive carbides have been
decomposed. Since graphitization is a relatively slow process, the casting must be soaked at
temperature for at least 20 hours and large loads may require as much as 72 hours. The structure
at completion of first stage graphitization consists of temper carbon nodules distributed throughout
the matrix of saturated austenite.
After first stage of annealing, castings are cooled as rapidly as practical to about 1400F in
preparation for the second stage of the annealing heat treatment. The fast cooling cycle usually
requires 2 to 6 hours depending on the equipment used.
In the second stage annealing, the castings are cooled slowly at a rate of 5 to 15F/h through the
critical range at which the eutectoid reaction will take place. During the slow cooling, the carbon
dissolved in the austenite is converted to graphite on the existing temper carbon particles, and the
remaining austenite transforms to ferrite. Once graphitization is complete, no further structural
changes take place during cooling to room temperature, and the structure consists of temper
carbon nodules in a ferrite matrix. This type is known as standard or ferritic malleable iron.
In the form of compact nodules, the temper carbon does not break up the continuity of the tough
ferritic matrix. This results in a higher strength and ductility than exhibited by gray cast iron. The
graphite nodules also serve to lubricate cutting tools, which account for the very high machinability
of malleable iron. Ferritic malleable iron has been widely used for automotive, agricultural and
railroad equipment, railing casting on bridges, chain hoist assemblies, pipe fittings and many
applications in general hardware.
Alloyed malleable irons are those whose properties result from the addition of alloying elements.
Since alloyed malleable irons are completely malleableized, their influence is largely on the ferritic
matrix. Two principal kinds are copper alloyed and copper-molybdenum alloyed malleable iron. The
effect of copper is to increase corrosion resistance, tensile strength and yield point at very slight
reduction in ductility. Hardness is also increased. The addition of copper and molybdenum in
combination produces a malleable iron of superior corrosion resistance and mechanical properties.
The strength and hardness of pearlitic malleable iron is more than ferritic malleable iron. In this
type of iron, the structure consists of temper carbon nodules in a pearlite matrix. For making
pearlitic malleable iron, a controlled quantity of carbon, in the order of 0.3 to 0.9 percent is
retained as finely distributed iron carbide during first stage of annealing.
If manganese is added, the regular cycle (of second annealing) can be maintained to retain
combined carbon throughout the matrix, or the second stage annealing of the normal process be
replaced by a quench, usually air, which cools the castings through the eutectoid range fast
enough to retain combined carbon throughout the matrix. The amount of pearlite formed depends
upon the temperature at which the quench starts and the rate of cooling. If the air quench
produces a fast enough cooling rate through the eutectoid range, the matrix will be completely
pearlitic. It is a common practice to temper most pearlitic malleable iron after air cooling to
spheroidize the pearlite, lower hardness, and improve machinability and toughness.
Pearlitic malleable irons are used for axle and differential housings, camshafts and crankshafts in
automobiles, gears, elevator buckets in conveyor equipment, pumps, nozzles, hammers, etc.
For comparison, tensile properties of different types of malleable iron are given in the following
table.

Type

Tensile strength,
1000 psi

Yield Strength,
1000 psi

Elongation,
% in 2 in.

Ferritic

50-60

32-39

20-10

110-145

Cu-Mo Alloyed Ferritic

58-65

40-45

20-15

135-155

Pearlitic

65-120

45-100

16-2

163-269

Gray Cast Iron

Cast irons are alloys of iron, carbon, and silicon in which more carbon is present than can be
retained in solid solution in austenite at the eutectic temperature. In gray cast iron, the carbon that
exceeds the solubility in austenite precipitates as flake graphite. Gray irons usually contain 2.5 to
4% C, 1 to 3% Si, and additions of manganese, depending on the desired microstructure (as low as
0.1% Mn in ferritic gray irons and as high as 1.2% in pearlitics). Sulphur and phosphorus are also
present in small amounts as residual impurities.
Gray cast iron is by far the oldest and most common form of cast iron. As a result, it is assumed by
many to be the only form of cast iron and the terms cast iron and gray iron are used
interchangeably. Unfortunately the only commonly known property of gray iron brittleness is
also assigned to cast iron and hence to all cast irons. Gray iron is named so because its fracture
has a gray appearance. It contains carbon in the form of flake graphite in a matrix which consists of
ferrite, pearlite or a mixture of the two. In the manufacture of gray cast irons, the tendency of
cementite to separate into graphite (graphitization) and austenite or ferrite/pearlite is aided by
high carbon content and the proper amount of graphitizing elements, notably silicon. With proper
control, the alloy will form austenite and graphite at the eutectic temperature (of the stable irongraphite equilibrium diagram) of 2075F.At any rate, any cementite which is formed will graphitize
rapidly. The graphite appears as many irregular, generally elongated and curved plates which give
gray cast iron its characteristic grayish or blackish fracture.
The strength of gray cast iron depends almost entirely on the matrix in which the graphite is
embedded. This matrix is largely determined by the condition of the eutectoid cementite. If the
composition and cooling rate are such that the eutectoid cementite also graphitizes, then the
matrix will be entirely ferritic. On the other hand, if graphitization of the eutectoid cementite is
prevented, the matrix will be entirely pearlitic. The constitution of the matrix may be varied from
pearlitic, through mixture of pearlitic and ferritic in different proportions, down to practically pure
ferritic. Moderate cooling rate results in a pearlitic gray iron where as slow cooling rate results in a

ferritic gray iron. The graphite-ferrite mixture is the softest and weakest gray iron. The strength and
hardness increases with the increase in combined carbon, reaching a maximum with the pearlitic
gray iron.
Silicon is very important element in the metallurgy of gray iron. It is a graphitizer and also
increases fluidity. Therefore, during solidification in the presence of silicon, carbon is precipitated as
primary graphite in the form of flakes. Once primary graphite has formed, its shape cannot be
altered by any method. It is these weak graphite flakes that break up the continuity of the matrix
and the notch effect at the end of these flakes that accounts for the low strength and low ductility
of gray iron.
Following figure shows percentage of carbon and silicon in cast irons and steels.

Heat Treatment of Gray Iron


Stress relieving is probably the most frequently applied heat treatment for gray irons. Gray iron in
the as-cast condition usually contains residual stresses because cooling proceeds at different rates
throughout various sections of a casting. The resultant residual stresses may reduce strength,
cause distortion, and in some cases even result in cracking.
Annealing of gray iron improves its machinability.
Gray iron is normalized to enhance its mechanical properties, such as hardness and tensile
strength, or to restore as-cast properties that have been modified by another heat treating process,
such as preheating and post heating associated with repair welding.
Gray iron, like steel, can be hardened and tempered. The high matrix hardness and graphite as
lubricant results in a surface with good wear resistance for some applications such as farm
implement gears, sprockets, diesel cylinder liners, etc.
Thus, heat treatment extends the field of application of gray iron as an engineering material.

Size and Distribution of Graphite Flakes


Large graphite flakes seriously interrupt the continuity of pearlitic matrix, thereby reducing the
strength and ductility of gray iron. Small graphite flakes are less damaging and are therefore
generally preferred.
Graphite-flak sizes are usually determined by comparison with standard sizes prepared jointly by
the AFS (American Foundrymens Society) and the ASTM (American Society for Testing
Materials). The procedure for preparation and measurement of flake size is given in ASTM
standards specification A247. The measurement is made of the lengths of the largest graphite
flakes in an unetched section of the gray iron at 100x. Numbers are assigned as indicated in the
table given below.

AFS-ASTM Flake Size Number

Length of Longest Flake at 100x, MM

128

64

32

16

Length of Longest Flake at 100x, inch

2-4

1-2

1/2-1

1/4-1/2

1/8-1/4

1/16-1/8

The way in which the graphite lakes are arranged in the microstructure of gray iron is usually
indicated as one or more types that have been jointly prepared by the AFS and the ASTM. The five
flake types are as under.

Type
Type
Type
Type
Type

A: Uniform distribution, random orientation


B: Rosette groupings, random orientation
C: Superimposed flake sizes, random orientation
D: Interdendritic segregation, random orientation
E: Interdendritic segregation, preferred orientation

Mechanical Properties of Gray Cast Iron


The most important classification of gray cast iron is given in the ASTM Specification A48. The gray
iron castings are classed in seven classes (Nos. 20, 25, 30, 35, 40, 50 and 60) which give the
minimum tensile strength of test bars in thousands of pounds per square inch. For example, class
20 gray iron would have a minimum tensile strength of 20000 psi.
Gray irons do not exhibit a well defined yield point as do mild steels. The percent elongation is
small for all cast irons, rarely exceeding 3 to 4 percent, and the reduction of area is too small to be
appreciable.

Many grades of gray iron have higher torsional shear strength than some grades of steel. This
characteristic, along with low notch sensitivity, makes gray iron a suitable material for various
types of shafting.

Applications of Gray Cast Iron


The fluidity of liquid gray iron, and its expansion during solidification due to the formation of
graphite, has made this metal ideal for the economical production of shrinkage-free, intricate
castings.
Because gray iron is the least expensive type of casting, it should always be considered first when
a cast metal is being selected. Another metal should be chosen only when the mechanical and
physical properties of gray iron are inadequate.
Examples of applications requiring bare minimum casting properties and lowest possible cost are
counterweights for elevators and industrial furnace doors. Gray iron castings are widely used for
gear housings, steam turbine housings, enclosures for electrical equipment, pump housings, motor
frames and sewer covers

Chilled Cast Iron


Chilled-iron castings are made by casting the molten metal against a metal chiller, resulting in a
surface of white cast iron. This hard, abrasion-resistant white-iron surface or case is backed up by a
softer gray iron core. This case-core structure is obtained by careful control of the overall alloy
composition and adjustment of the cooling rate.
Freezing starts first, and cooling rate is most rapid where the molten metal is in contact with the
mold walls. The cooling rate decreases as the center of the casting is approached.
If only selected surfaces are to be white iron, it is common practice to use a composition which
would normally solidify as gray iron and employ metal liner (chills) to accelerate the cooling rate of
the selected areas. The depth of the white iron layer is controlled by using thin metal plates
whenever a thin white iron layer is desired and heavier metal plates where a deeper chill is
necessary.
Chilled iron casting is used for railway-car wheels, crushing rolls, sprockets, and many other heavyduty machinery parts.

Nodular Cast Iron

Nodular cast iron, also known as ductile iron (in UAS), spheroidal graphite iron (in UK and Europe)
and spherulitic iron is cast iron in which the graphite is present as tiny balls or spheroids. The
compact spheroids interrupt the continuity of the matrix much less than graphite flakes, resulting
in higher strength and toughness compared with a similar structure of gray cast iron. Nodular cast

iron differs from malleable iron in that it is usually obtained as a result of solidification and does not
require heat treatment. The spheroids are more rounded than the irregular aggregates of temper
carbon found in malleable iron.
The total carbon content of nodular iron is the same as in gray iron. Spheroidal graphite particles
form during solidification because of the presence of a small amount of certain alloying elements.
The nodule forming addition, usually magnesium or cerium is made to the ladle just before casting.
Since these elements have a strong affinity for sulfur, in the base iron-alloy, sulfur content must be
below 0.015 percent for the treatment to be effective, and the alloys are described as desulfurized.
ASTM A536 Standard Specification for Ductile Iron Castings is the most frequently used
specification, covering the general engineering grades of ductile iron.
Standard specifications for ductile iron are normally based on mechanical properties, except for
those defining austenitic ductile iron, which are based on composition. Mechanical requirements for
various grades of ductile irons as per ASTM A536 are as under.

Property

Grade 60/40/18

Grade 65/45/12

Grade 80/55/06

Grade 100/70/03 Grade

Tensile strength, min, psi

60 000

65 000

80 000

100 000

120 00

Tensile strength, min, MPa

414

448

552

689

827

Yield strength, min, psi

40 000

45 000

55 000

70 000

90 000

Yield strength, min, MPa

276

310

379

483

621

Elongation in 2 in. min, %

18

12

6.0

3.0

2.0

The amount of ferrite in the as-cast matrix depends on composition and rate of cooling. Nodular
iron with a matrix having a maximum of 10 percent pearlite is known as ferritic irons. This structure
gives maximum ductility, toughness and machinability.
A matrix structure which is largely pearlitic can be produced as cast or by normalizing. Pearlitic
ductile irons are stronger but less ductile than ferritic iron.
A martensitic matrix may be obtained by quenching in oil or water. The quenched structures are
usually tempered after hardening to desired strength and hardness.
Austenitic ductile irons are highly alloyed types which retain their austenitic structure down to at
least minus 75F. They have relatively high corrosion resistance and good creep properties at
elevated temperatures. More information of them is given in this article under alloy cast irons.
Tensile mechanical properties of basic types of nodular iron are given in the following table.

Type

Alloy Content

Tensile Strength,
psi

Yield Strength,
psi

Elongation,
% in 2 in.

Low

55000

35000

25

130

High

90000

70000

12

210

Low

80000

60000

10

200

Ferritic

Pearlitic

Low (Normalized)

130000

90000

275

High

130000

110000

275

100000

80000

10

215

150000

130000

320

60000

30000

40

130

**

60000

40000

10

160

Quenched

Austenitic

* 3.00 % C, 2.50 % Si, 20.00 % Ni, 2.00 % Mn


** 3.00 % C, 2.00 % Si, 20.00 % Ni, 1.00 % Mn, 1.50 % Cr
Nodular iron is used in agricultural tractor and implement parts, automotive and diesel
crankshafts, pistons and cylinder heads, electrical fittings, motor frames, circuit breaker parts,
Mining hoist drums, drive pulleys, flywheels and elevator buckets, steel mills, tools and dies.

Alloy Cast Irons


An alloy cast iron is one which contains a specially added element or elements in sufficient amount
to produce a measurable modification in the physical or mechanical properties.
Alloying elements are added to cast iron for special properties such as resistance to corrosion, heat
or wear and to improve mechanical properties. The most common alloying elements are nickel,
chromium, molybdenum and copper. For various industrial applications requiring resistance against
wear, corrosion and heat, two types of alloy cast irons, High-alloy white cast irons and Ni-Resist
cast irons are widely used. Information about both these varieties is given in the sections that
follow.

High-alloy White Cast Irons


The high-alloy white irons are primarily used for abrasion-resistant applications. The chromium
content of high-alloy white irons also enhances their corrosion-resistant properties. The large
volume fraction of primary and/or eutectic carbides in their microstructures provides the high
hardness needed for crushing and grinding other materials. The metallic matrix supporting the
carbide phase in these irons can be adjusted by alloy content and heat treatment to develop the
proper balance between the resistance to abrasion and the toughness needed to withstand
repeated impact.
While low-alloy white iron castings, which have alloy content below 4%, develop hardness in the
range of 350 to 550 HB, the high-alloy irons range in hardness from 450 to 800 HB.
Specification ASTM A532 covers the composition and hardness of the abrasion-resistant white iron
grades. This specification deals with abrasion-resistant cast irons used for mining, milling, earthhandling, and manufacturing industries. It may be noted that simple and low-alloy white cast irons
that consist essentially of iron carbides and pearlite are specifically excluded from this
specification.
The high-alloy white cast irons fall into two major groups

Nickel-chromium white irons


They are low-chromium alloys containing 3 to 5% Ni and 1 to 4% Cr, with one alloy modification
that contains 7 to 11% Cr. The nickel-chromium irons are also commonly identified as Ni-Hard types
1 to 4.

Chromium-molybdenum irons
They contain 11 to 23% Cr, up to 3.5% Mo and often additionally alloyed with nickel or copper.

A third group comprises the 25% or 28% Cr white irons, which may contain other alloying additions
of molybdenum and/or nickel up to 1.5%.
The chemical composition for various class and type of high-alloy white irons as per ASTM A532 in
% is as under.

Specification

Mn

Si

Cr

Ni

Mo

Cu

ASTM A532 I-A

3.0-3.6

1.3

0.8

1.4 -4.0

3.3-5.0

1.0

0.15

ASTM A532 I-B

2.5-3.0

1.3

0.8

1.4 -4.0

3.3-5.0

1.0

0.15

ASTM A532 I-C

2.9-3.7

1.3

0.8

1.1 -1.5

2.7 -4.0

1.0

0.15

ASTM A532 I-D

2.5-3.6

1.3

1.0 -2.2

7.0 11.0

4.5-7.0

1.0

0.15

ASTM A532 II-A

2.4-2.8

0.5-1.5

1.0

11.0-14.0

0.5

0.5-1.0

1.2

0.06

ASTM A532 II-B

2.4-2.8

0.5-1.5

1.0

14.0-18.0

0.5

1.0-3.0

1.2

0.06

ASTM A532 II-C

2.8-3.6

0.5-1.5

1.0

14.0-18.0

0.5

2.3-3.5

1.2

0.06

ASTM A532 II-D

2.0-2.6

0.5-1.5

1.0

18.0-23.0

1.5

1.5

1.2

0.06

ASTM A532 II-E

2.6-3.2

0.5-1.5

1.0

18.0-23.0

1.5

1.0-2.0

1.2

0.06

ASTM A532 III-A

2.3-3.0

0.5-1.5

1.0

23.0-28.0

1.5

1.5

1.2

0.06

Nickel-Chromium White Irons


In these martensitic white irons, nickel is the primary alloying element because at levels of 3 to 5%
it is effective in suppressing the transformation of the austenite matrix to pearlite, thus ensuring
that a hard martensitic structure (usually containing significant amounts of retained austenite) will
develop upon cooling in the mold. Chromium is included in these alloys, at levels from 1.4 to 4%, to
ensure that the irons solidify carbidic, that is, to counteract the graphitizing effect of nickel. These
types of white irons are used for crushing and grinding applications.
The optimum composition of a nickel-chromium white iron alloy depends on the properties required
for the service conditions and the dimensions and weight of the casting. Abrasion resistance is
generally a function of the bulk hardness and the volume of carbide in the microstructure. When
abrasion resistance is the principal requirement and resistance to impact loading is secondary,
alloys having high carbon contents, ASTM A532 class I type A (Ni-Hard 1), are recommended. When
conditions of repeated impact are anticipated, the lower carbon alloys, ASTM A532 class I type B
(Ni-Hard 2) are recommended because they have less carbide and, therefore, greater toughness. A
special grade, class I type C, has been developed for producing grinding balls and slugs. Here, the
nickel-chromium alloy composition has been adapted for chill casting and specialized sand casting
processes.
The ASTM A532 Class I type D (Ni-Hard 4) alloy is a modified nickel-chromium iron that contains
higher levels of chromium, ranging from 7 to 11%, and increased levels of nickel, ranging from 5 to
7%. Carbon is varied according to the properties needed for the intended service. Carbon contents
in the range of 3.2 to 3.6% are prescribed when maximum abrasion resistance is desired. When
impact loading is expected, carbon content should be held in the range of 2.7 to 3.2%.
Nickel content increases with section size or cooling time of the casting to inhibit pearlitic
transformation. For castings of 38 to 50 mm thick, 3.4 to 4.2% Ni is sufficient to suppress pearlite
formation upon mold cooling. Heavier sections may require nickel levels up to 5.5% to avoid the

formation of pearlite. It is important to limit nickel content to the level needed for control of
pearlite; excess nickel increases the amount of retained austenite and lowers hardness.
Silicon is needed for two reasons. A minimum amount of silicon is necessary to improve fluidity of
the melt and to produce a fluid slag, but of equal importance is its effect on as-cast hardness.
Increased levels of silicon, in the range of 1 to 1.5%, have been found to increase the amount of
martensite and the resulting hardness. It is important to note that higher silicon contents can
promote pearlite and may increase the nickel requirement.
Chromium is primarily added to offset the graphitizing effects of nickel and silicon in types A, B,
and C alloys, ranges from 1.4 to 3.5%. Chromium content must be increased with increasing
section size. In type D alloy, chromium levels range from 7 to 11% (typically 9%) for the purpose of
producing eutectic carbides of the M7C3 chromium carbide type, which are harder and less
deleterious to toughness.
Manganese is typically held to a maximum of 0.8% even though 1.3% maximum is allowed
according to ASTM A532 specification. While it provides increased hardenability to avoid pearlite
formation, it is a more potent austenite stabilizer than nickel, and promotes increased amounts of
retained austenite and lower as-cast hardness. For this reason, higher manganese levels are
undesirable. When considering the nickel content required to avoid pearlite in a given casting, the
level of manganese present should be a factor.
Copper increases both hardenability and the retention of austenite and therefore must be
controlled for the same reason that manganese must be limited. Copper should be treated as a
nickel substitute and, when properly included in the calculation of the amount of nickel required to
inhibit pearlite, it reduces the nickel requirement.
Molybdenum is a potent hardenability agent in these alloys and is used in heavy-section castings to
augment hardenability and inhibit pearlite.

High-Chromium White Irons


The high-chromium white irons have excellent abrasion resistance and are used effectively in slurry
pumps, coal-grinding mills, shot-blasting equipment, and components for quarrying, hard-rock
mining, and milling. In some applications they must also be able to withstand heavy impact
loading. These alloyed white irons are recognized as providing the best combination of toughness
and abrasion resistance attainable among the white cast irons.
In the high-chromium irons, as with most abrasion-resistant materials, there is a trade-off between
wear resistance and toughness. By varying composition and heat treatment, these properties can
be adjusted to meet the needs of most abrasive applications.
Specification ASTM A532 covers the compositions and hardnesses of two general classes of the
high-chromium irons. The chromium-molybdenum irons (Class II of ASTM A532) contain 11 to 23%
Cr and up to 3.5% Mo and can be supplied either as-cast with an austenitic or austeniticmartensitic matrix, or heat-treated with a martensitic matrix microstructure for maximum abrasion
resistance and toughness. They are usually considered the hardest of all grades of white cast irons.
Compared to the lower-alloy nickel-chromium white irons, the eutectic carbides are harder and can
be heat-treated to achieve castings of higher hardness. Molybdenum, as well as nickel and copper
when needed, is added to prevent pearlite and to ensure maximum hardness.
The high-chromium irons (class III of ASTM A532) represent the oldest grade of high-chromium
irons. These general-purpose irons, also called 25% Cr and 28% Cr irons, contain 23 to 28% Cr with
up to 1.5% Mo. To prevent pearlite and attain maximum hardness, molybdenum is added in all but
the lightest-cast sections. Alloying with nickel and copper up to 1% is also practiced. Although the
maximum attainable hardness is not as high as in the class II chromium-molybdenum white irons,
these alloys are selected when resistance to corrosion is also desired.

Ni-Resist Cast Irons


Austenitic gray iron castings known as Ni-Resist irons (also called standard Ni-Resist irons) and
austenitic ductile iron castings known as ductile Ni-Resist irons are used primarily for their
resistance to heat, corrosion, and wear. Both families of alloys have similar general characteristics
as described below.

General characteristics of the Ni-Resist irons


Corrosion Resistance:
They have been specified to solve corrosion problems involving the handling of sour well oils, salts,
salt water, acids and alkalies. They are intermediate in corrosion resistance between gray cast
irons and austenitic chromium-nickel stainless steels. The excellent erosion-corrosion resistance of
the Ni-Resist alloys has resulted in extensive applications for pump and valve components in sea
water handling systems.
Wear Resistance:
Engine, pump and their parts like pistons, wearing rings, sleeves, glands, etc. and other metal-tometal rubbing parts have been cast in the Ni-Resist alloys.
Erosion Resistance:
Slurries, wet steam and gases with entrained particles are among the elements extremely erosive
to most metals. The Ni-Resist alloys offer a combination of corrosion-erosion-resistant properties
that provide superior service under many erosive conditions.
Toughness and Low-Temperature Stability:
Ni-Resist castings are considerably superior to gray iron under these service conditions.
Controlled Expansion:
Expansivities from 2.2 to 10.6 x 10-6 in./in. per degree F are available with different types of NiResist irons. This affords a conventional cast metal with low expansivity for precision machines or
parts.
Heat Resistance:
For unusual high-temperature service, 1600F and above, the ductile grade of Ni-Resist should be
considered. The standard grades (Ni-Resist irons) exhibit a high order of heat resistance to
1300F. The alloys have a relatively low rate of oxidation, and what oxidation does occur adheres
tenaciously to the base metal. This property is especially helpful in gas turbine, manifold and
turbocharger applications.
Castability:
The Castability of Ni-Resist permits casting complicated and intricate designs often difficult to
produce with some other high-alloy cast metals. Both the castability and machinability of the NiResist alloys aid in producing finished parts economically.

Ni-Resist Irons (Austenitic Gray Iron Castings)


Basically, the family is a series of cast irons to which has been added sufficient nickel to produce
an austenitic structure similar to that of austenitic stainless steel. Austenitic gray iron is
characterized by uniformly distributed graphite flakes, some carbides, and an austenitic structure.
This structure provides the alloy familys heat and corrosion resistant properties.
The several types of Ni-Resist are produced by varying the nickel content and, to some extent, the
chromium content and are covered by ASTM A436 Standard Specification for Austenitic Gray Iron
Castings. The types of castings covered are Type 1, Type 1b, Type 2, Type 2b, Type 3, Type 4, Type
5, and Type 6. Their chemical composition in % is as under.

Specification

Mn

Si

Cr

Ni

Mo

Cu

ASTM A436 TYPE1

3.00

0.5-1.5

1.00-2.80

1.50-2.50

13.50-17.50

5.50-7.50

ASTM A436 TYPE 1b

3.00

0.5-1.5

1.00-2.80

2.50-3.50

13.50-17.50

5.50-7.50

ASTM A436 TYPE 2

3.00

0.5-1.5

1.00-2.80

1.5 -2.5

18.00-22.00

0.50

ASTM A436 TYPE 2b

3.00

0.5-1.5

1.00-2.80

3.00-6.00

18.00-22.00

0.50

ASTM A436 TYPE 3

2.60

0.5-1.5

1.00-2.00

2.50-3.50

28.00-32.00

0.50

ASTM A436 TYPE 4

2.60

0.5-1.5

5.00-6.00

4.50-5.50

29.00-32.00

0.50

ASTM A436 TYPE 5

2.40

0.5-1.5

1.00-2.00

0.10

34.00-36.00

0.50

ASTM A436 TYPE 6

3.00

0.5-1.5

1.50-2.50

1.00-2.00

18.00-22.00

1.00

3.5 -5.5

The low nickel content of Types 1 and 1b is accompanied by copper to provide the corrosionresisting characteristics. (Note: Types 1 and 1b are not responsive to the magnesium treatment
process, and have no ductile counterparts.)
Mechanical requirements for various types of Ni-Resist irons covered by ASTM A436 are as under.

Type
Property
1

1b

2b

Tensile strength, min, ksi

25

30

25

30

25

25

Brinell hardness

131-183

149-212

118-174

171-248

118-159

149-212

Characteristics and applications of the different types are as under.


Types 1 and 2: are used interchangeably in many corrosion and wear resistant applications. Both
are specially suited to heavy metal-to-metal wear service.
Type 1: has some advantages over the other types in handling mineral-acid corrosives and salt
water.
Type 1b: has superior corrosion-erosion resistance in comparison with Type 1, and is harder and
stronger.
Type 2: because of its wide use in corrosive environments, is the most commonly used. It is
preferred for heat and oxidation resistance to 1300F and in steam service. It is also used for
handling caustics, alkalies, ammoniacal solutions, food products, rayon, plastics and similar
environments where freedom from copper contamination is desired.
Type 2b: is especially recommended for heat applications up to 1500F. Uses include gas turbine
parts, exhaust manifolds and turbochargers. If machining is required, the chromium level should be
kept between 3.0 and 4.0 percent. The greater hardness of Type 2b makes it suitable for resistance
to abrasive wear with corrosion, but is not suited for metal-to-metal wear.
Type 3: is recommended for sever thermal shock service between room temperature and 450F.
At temperatures between 450 and 1500F it can be used without severe thermal shock. It may be
used for diesel exhaust manifolds and turbochargers. The type offers high resistance to erosion in
wet steam and corrosive slurries.
Type 4: is recommended where strain-resistance is required. This type is superior to other types in
resistance to erosion, corrosion and oxidation.
Type 5: offers minimum thermal expansivity which provides dimensional stability for machine tool
parts, forming dies, steam turbines, scientific instruments and expansion joints.

Ductile Ni-Resist Irons (Austenitic Ductile Iron Castings)


The ductile Ni-Resist irons are similar to standard/conventional Ni-Resist compositions but have
been treated with magnesium to convert the graphite from the flake form to spheroids, which
results in higher strength and ductility. Austenitic ductile iron, also known as austenitic nodular iron
or austenitic spheroidal iron, is characterized by having its graphite substantially in a spheroidal
form and substantially free of flake graphite. It contains some carbides and sufficient alloy content
to produce an austenitic structure.
The ductile family of Ni-Resist is available in every type except Type 1, which because of the higher
copper content does not respond properly to the magnesium treatment. There are in addition
several modifications of Type 2 to 5. Various types of ductile Ni-Resist irons are covered by ASTM
A439 Standard Specification for Austenitic Ductile Iron Castings. Their chemical composition in %
is as under.

Specification

Mn

Si

Cr

Ni

ASTM A439 Type D-2

3.00

0.70-1.25

1.50-3.00

1.75-2.75

18.00-22.00

0.0

ASTM A439 Type D-2B

3.00

0.70-1.25

1.50-3.00

2.75-4.00

18.00-22.00

0.0

ASTM A439 Type D-2C

2.90

1.80-2.40

1.00-3.00

0.50

21.00-24.00

0.0

ASTM A439 Type D-3

2.60

1.00

1.00-2.80

2.50-3.50

28.00-32.00

0.0

ASTM A439 Type D-3A

2.60

1.00

1.00-2.80

1.00-1.50

28.00-32.00

0.0

ASTM A439 Type D-4

2.60

1.00

5.00-6.00

4.50-5.50

28.00-32.00

0.0

ASTM A439 Type D-5

2.40

1.00

1.00-2.80

0.10

34.00-36.00

0.0

ASTM A439 Type D-5B

2.40

1.00

1.00-2.80

2.00-3.00

34.00-36.00

0.0

ASTM A439 Type D-5S

2.30

1.00

4.90-5.50

1.75-2.25

34.00-37.00

0.0

Mechanical requirements for various types of ductile Ni-Resist irons covered by ASTM A439 are as
under.

Type
Property
D-2

D-2B

D-2C

D-3

D-3A

D-4

D-5

D-5B

Tensile strength, min, ksi (MPa)

58 (400) 58 (400) 58 (400) 55 (379) 55 (379) 60 (414) 55 (379) 55 (379

Yield strength (0.2 percent offset),


min, ksi (MPa)

30 (207) 30 (207) 28 (193) 30 (207) 30 (207) -

30 (207) 30 (201

Elongation 2 in. or 50mm, min, %

8.0

20.0

Brinell hardness (300 kg)

139-202 148-211 121-171 139-202 131-193 202-273 131-185 139-193

7.0

20.0

6.0

10.0

6.0

Characteristics and applications of the different types are as under.


Type D-2: is recommended for service requiring resistance to corrosion, erosion, frictional wear and
temperature up to 1400F. This type also has high thermal expansivity.
Type D-2B: is recommended for superior resistance to neutral and reducing salts, and where higher
resistance to erosion and oxidation than that offered by Type D-2 is desired.
Type D-2C: is recommended where resistance to heat and corrosion is less severe and where high
ductility is desired.
Type D-3: is recommended for thermal shock service and where the thermal expansivity should
match that of ferritic stainless steels. This type, in addition to having excellent elevatedtemperature properties, also offers high resistance to erosion.
Type D-3A: is recommended for use where a high degree of wear and galling resistance is required
along with intermediate thermal expansion.
Type D-4: is superior to Types D-2 or D-3 in resistance to corrosion, erosion and oxidation.
Type D-5: is utilized whenever minimum thermal expansion is desired. It may also be preferred over
other types of Ni-Resist to reduce thermal stress.
Type D-5B: is recommended for applications requiring a very low order thermal stress. In addition,
this type has a high level of heat and oxidation resistance as well as good mechanical properties at
elevated temperatures.
Type D-5S: is recommended where oxidation resistance in air to 1800F is desired. It also resists
embrittlement and offers thermal stability and strength in cyclic heating to 1600F.

Summary of the cooling rates to make various types of cast


irons
A summary of the cooling rates to make various types of cast iron and their microstructure at room
temperature is given below.

S-ar putea să vă placă și