Você está na página 1de 12

400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-5760 Web: www.sae.

org
SAE TECHNICAL
PAPER SERIES
2004-01-0792
Solidification Behavior, Microstructure,
Mechanical Properties, Hot Oxidation
and Thermal Fatigue Resistance of
High Silicon SiMo Nodular Cast Irons
D. Li, R. Perrin, G. Burger, D. McFarlan,
B. Black, R. Logan and R. Williams
Wescast Industries Inc.
Reprinted From: Advances in Lightweight Automotive Castings
and Wrought Aluminum Alloys
(SP-1838)
2004 SAE World Congress
Detroit, Michigan
March 8-11, 2004
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise,
without the prior written permission of SAE.
For permission and licensing requests contact:
SAE Permissions
400 Commonwealth Drive
Warrendale, PA 15096-0001-USA
Email: permissions@sae.org
Fax: 724-772-4891
Tel: 724-772-4028
For multiple print copies contact:
SAE Customer Service
Tel: 877-606-7323 (inside USA and Canada)
Tel: 724-776-4970 (outside USA)
Fax: 724-776-1615
Email: CustomerService@sae.org
ISBN 0-7680-1319-4
Copyright 2004 SAE International
Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE.
The author is solely responsible for the content of the paper. A process is available by which discussions
will be printed with the paper if it is published in SAE Transactions.
Persons wishing to submit papers to be considered for presentation or publication by SAE should send the
manuscript or a 300 word abstract of a proposed manuscript to: Secretary, Engineering Meetings Board, SAE.
Printed in USA
2004-01-0792
Solidification Behavior, Microstructure, Mechanical Properties,
Hot Oxidation and Thermal Fatigue Resistance of
High Silicon SiMo Nodular Cast Irons
D. Li, R. Perrin, G. Burger, D. McFarlan, B. Black, R. Logan and R. Williams
Wescast Industries Inc.
Copyright 2004 SAE International
ABSTRACT
It is well known that 4 to 6% silicon spheroidal irons are
suitable for use at high temperature. This paper describes
solidification behavior, microstructure, mechanical properties,
high temperature oxidation, and thermal fatigue of high silicon
SiMo cast irons. Cooling curves of cast irons were recorded
using a thermal analysis apparatus to correlate with the
solidified microstructures. Uniaxial constrained thermal
fatigue testing was conducted in which the cycling
temperatures were between 500
o
C and 950
o
C. Oxidation
behavior was studied by measuring the specimen weight and
the penetration depth of oxides from laboratory cyclic
oxidation testing. The coefficient of thermal expansion and
critical temperature of the phase transformation A1 during
heating were determined through dilatometry testing.
INTRODUCTION
The high-silicon (high-Si) and silicon-molybdenum (SiMo)
ductile irons are currently used to produce exhaust manifolds
and turbocharger housings for automobiles. With increasing
Si content, the high temperature oxidation/spallation
resistance, the critical temperature A1, and microstructural
stability are enhanced [1-4]. Studies [5] have also shown that,
to a significant extent, Si increases uniaxial thermal fatigue
life and can, to a certain extent, replace Mo. The drawbacks to
increased Si levels include a lower impact toughness and
higher Brinell hardness, which may possibly cause some
casting handling and machining problems during mass-
production. Also certain manifold designs can be susceptible
to thermal cracking when utilizing this material with too high
Si contents. This may be due to lower overall ductility and
thermal conductivity of this material compared to some other
manifold irons. Typical SiMo ductile irons containing 4% Si
and 0.5 to 2% Mo [6] can be used up to the material
temperatures of 840 to 850
o
C [7] and in some cases with
beneficial designs, even a little higher. The primary objective
of this work is to examine if SiMo irons with a higher Si
content ranging from 4.4 to 5.0% can be used for large-scale
production of exhaust manifolds. Increased Si additions will
change the freezing behavior, shrinkage tendency, elongation,
impact toughness, oxidation resistance, thermal fatigue life,
casting handling, and machinability. This work was
undertaken to investigate the suitability of high-Si SiMo for
exhaust manifold applications, and to document the properties
mentioned above through both experiments and modeling.

EXPERIMENTAL PROCEDURES
A number of regression trials and DOE (design of
experiments) on alloy chemistry and processing were
performed at the Richard W. Levan Technical Centre of
Wescast Industries Inc. Heats weighing approximately 100 kg
were made using an induction furnace of 350 kW and 1 kHz.
Conventional treatment and inoculation were used. Pour
temperatures varied from 1390
o
C to 1460
o
C. Test bars and
prototype manifolds were cast in chemically bonded sand and
greensand molds respectively. Cooling curves of cast iron
melts were measured using a thermal analysis apparatus.
Approximately 240 manifolds with high-Si SiMo were cast
and machined during the plant trials to evaluate handling and
machinability.

Optical microscopy and digital image analysis software were
used for completing the microstructural analysis of the trial
samples. In addition to casting experiments, solidification
modeling was conducted for different chemistries and pour
temperatures. The solid phase transformation temperature and
coefficient of thermal expansion were determined by
dilatometry testing at the heating rate of 1 and 5
o
C/min
separately. Muffle furnaces were utilized for elevated
temperature oxidation testing. The oxidation specimens were
101030 mm rectangular bars. The total surface area (1400
mm
2
) is larger than the minimum value (400 mm
2
)
recommended by ASTM G54-84. Before testing, samples
were cleaned ultrasonically and heat treated at 210
o
C for 2
hours. After obtaining the initial weight of each sample, the
oxidation testing started at 900
o
C. During testing the
specimens were withdrawn from the furnace, cooled, and
replaced at intervals of 24 to 48 hours. Room and elevated
temperature tensile and charpy impact testing were performed
at both internal and external laboratories. Y-blocks were cast
at Wescast Industries Inc. and provided to Climax Research
Services (CRS) in Michigan for uniaxial and constrained
thermal fatigue testing between 500
o
C and 950
o
C. The
specimens were heated by a radio-frequency (RF) furnace
(450 kHz) within a rigid test frame. Further descriptions of
the CRS facility and testing procedures have been given
elsewhere [5].

RESULTS AND DISCUSSION
SOLIDIFICATION AND SHRINKAGE - Understanding of
the solidification process is essential to control microstructure,
which in turn, determines the properties of materials. SiMo
cast irons can be considered as a quarternary Fe-C-Si-Mo
system. As the silicon content is increased, the carbon content
of the eutectic and eutectoid is decreased. Also the eutectic
and eutectoid temperatures change from a single value to a
temperature range. Thus a simple linear relation CE =
%C+1/3%Si is defined to represent the combined effect of
silicon and carbon, namely the carbon equivalent. A
chemistry map was drawn regarding C, Si, and CE with a
microstructure prediction, as presented in Figure 1. High-Si
SiMo discussed in this paper contains 4.4 to 5.0% Si and 0.5
to 0.9% Mo in iron, which is further divided into two groups
according to Si content: high-Si I (4.4 to 4.7% Si) and high-Si
II (4.7 to 5% Si). It is widely accepted that the CE values
should be controlled around 4.7 for castings of wall thickness
less than 25 mm in order to avoid solidification of austenite
dendrites and primary carbides. Too low or too high CE will
give rise to casting defects. This map can be further
elucidated by experimental results as depicted in Figure 2. For
a strongly hypereutectic chemistry such as CE = 4.89, the
primary liquidus thermal arrest is clearly visible, which was
referred to as the graphite liquidus [8], as illustrated in the
cooling curve (a) of Figure 2. This corresponds to the
formation of coarse graphite nodules prior to the eutectic
reaction, as shown in micrograph (b) of Figure 2. Referring to
curve (c) of Figure 2, the sample with CE = 4.7 shows one
solidification event, namely that the liquidus temperature and
the starting eutectic temperature are identical. The curve
exhibits fairly smooth cooling to the low eutectic temperature
where the bulk growth started followed by a small
recalescence of less than 10
o
C. This curve led to a uniform
nodule distribution, as seen in micrograph (d) of Figure 2.
The Si content was 4.95% in Figure 2. However, if the CE is
too low, both shrinkage and chilling tendency will increase.
Furthermore thermal conductivity of cast iron will decrease if
the C content is too low or/and the Si content is too high,
which may reduce the thermal fatigue resistance of castings
[1]. With regard to the alloying element Mo, part of it
segregated and froze into intercellular regions to promote
eutectic (lamellar shape) or primary (faceted shape) carbides.
Furthermore, during the solid state transformation, fine Mo-
rich particles of less than 1 m in diameter were precipitated
around the grain boundaries, as observed by Black et al. [9].
The degree of undercooling and the solidius temperature
determined from cooling curves may be useful to predict
formation of the primary carbides.

Solidification behavior and nodule distribution significantly
influences the shrinkage tendency. Figure 3 demonstrates the
relationship between the carbon equivalent and shrinkage
tendency through a series of studies including manifold
castings (the two pictures in the left side), AFS (American
Foundry Society) blind risers (the two pictures in the middle),
and a solidification software modeling (the right side column).
The CE value was 4.85 for the top row and 4.65 for the bottom
row of pictures; while the other conditions are basically the
same including the Si content (4.50%), mold materials,
magnesium treatment, inoculation, and pour temperature.
Simply speaking, the top samples produced much more severe
shrinkage than the bottom row. When the CE is too high,
solidification started with a higher liquidus temperature, thus
forming the mushy liquid earlier and macro shrinkage, as
shown in the top pictures of Figure 3. One can also say that
high shrinkage level melts were always formed with a large
number of coarser nodules [10]. For the bottom row with a
CE = 4.65, much less shrinkage was observed within the cross
sections of casting manifolds, AFS blind risers, and
solidification modeling pictures. Generally speaking,
shrinkage is also influenced by other many factors.

MECHANICAL PROPERTIES The silicon effects on
mechanical properties have been well studied and understood
[1-4]. When the Si content is lower than approximately 3.5%
[11], the tensile strength decreases and elongation increases
with increasing Si content, since more ferritic phase was
formed within the matrix. If Si content exceeds that value, the
tensile strength and hardness increase, while the elongation
and impact toughness decrease with increasing Si. In this
work, the Si content was higher than the critical value, and Mo
content varied from 0.5 to 0.9%. Y-blocks with different Si
contents were cast for various types of testing. Figure 4
presents the room temperature tensile and hardness data. It is
seen that with increasing Si content, the ultimate tensile
strength (UTS), 0.2% offset yield strength (YS), and the
Brinell hardness went up, due to the solid solution
strengthening and hardening, while the tensile elongation
dropped. However, elongation of up 10% can be steadily
achieved for 4.95% Si samples through controlling melt
chemistry and inoculation practice.

Standard short-time tensile testing at elevated temperatures
was conducted. Figure 5 presents some UTS data obtained
from our trials. It was observed that the Si strengthening
effect diminished when the testing temperature was beyond
500
o
C. On the contrary, the UTS at 700
o
C and 900
o
C slightly
decreased with increasing Si, perhaps because silicon reduced
the solubility of carbon in cast irons and tended to dissociate
iron carbides. More data are needed to examine whether
silicon really lowers the elevated-temperature strength or not.
It is well known that Mo can effectively increase the elevated
temperature strength, both short-time tensile strength and
long-term creep strength [1, 6].

The ductility is indeed reduced with increasing Si contents. A
natural question arises: Can heat treatment improve the
ductility of high-Si SiMo? Both sub-critical (785
o
C) and full
(950
o
C) annealing were conducted for high-Si SiMo test bars
of 16 mm in thickness. The samples were held at the
mentioned temperatures for 4 hours and then slowly cooled
(approximately 100
o
C/hr). Samples were removed from the
furnace for air cooling after reaching 600
o
C to avoid possible
elevated temperature embrittlement. As expected, the UTS
and YS slightly decreased and elongation increased after
annealing, as illustrated in Figure 6. Regarding the
microstructure of Mo containing ductile irons, there exists the
so-called Mo-rich phase around the cell-boundary regions [9].
The percentage of Mo-rich phase (Mp) can be determined
from image analysis of the microstructure. It was found that
the Mp content was reduced from 7.3% (as-cast), to 6.4%
(subcritical annealing), and to 3.2% (full annealing), for the
same chemistry (4.65% Si and 0.75% Mo). Less Mp
percentage increased the elongation. However, there was no
major change in the impact toughness before and after heat
treatment, as seen in Figure 7. The absorbed energy was
determined from the charpy impact tests which were
performed on the standard unnotched rectangular (101055
mm) bars. The samples with 4.65% Si show higher impact
toughness than those with 5.01% Si even with a large scatter
in the data. The unnotched charpy data of regular SiMo (4%
Si) ranged from 30 to 45 J which is much higher than high-Si
SiMo. The impact toughness is a measure of the amount of
energy a material can absorb before fracturing. The charpy
impact data indicate the handling severity of casting manifolds
in the foundry process including shakeout, media drum
cycling, and rotary shot-blaster. Plant trials with high-Si SiMo
chemistry were conducted in order to evaluate the possible
issues caused by rough handling and machining in large-scale
production.

HIGH TEMPERATURE OXIDATION - There are numerous
publications on high-temperature oxidation and scaling with
different alloy chemistries, testing temperatures and exposure
times [1-4, 12-14]. Oxidation and spallation may possibly
have detrimental effects such as reducing the stress-bearing
ability and causing particles to travel into downstream
component. In this paper, we report the results of SiMo with
4.0% Si and high-Si SiMo with 4.8% Si obtained from two
conditions respectively. No airflow was introduced into the
furnace under condition I, while under condition II fresh air
was passed through the furnace throughout the run at the rate
of 2.5 liters/min monitored by a flow meter. Under both
conditions, the furnace temperature was calibrated to 900
o
C,
and the samples were removed from the furnace and cooled
down to room temperature every 1 or 2 days (cyclic oxidation
testing). More than 20 cycles were run. Figure 8 shows the
accumulative weight change, but the weight change for each
period can be determined too from the curves. Under
condition I (empty symbols) the total weight gain was almost
nil for high-Si SiMo, while it was up 7 mg/cm
2
for SiMo.
Also SiMo shows much higher oxidation rate initially.
However, after 100 hours, the two curves (empty symbols)
have essentially the same slope. Under condition II (solid
symbols), the superiority of high-Si SiMo versus SiMo
becomes more evident. After 695 hours, the weight gain was
20.5 mg/cm
2
for SiMo, while it was only 6.1 mg/cm
2
for high-
Si SiMo. Furthermore, a steeper slope of the SiMo curve
(solid squares) indicates the degree of rapid oxidation.

The solid penetration depth can be considered as a more
meaningful measure of oxidation, because the weight change
in Figure 8 represents the combined effects of weight gained
by the formation of oxides and the weight lost by
decarburization. The weight loss was dominant during the
initial stage of the reaction. As oxidation proceeds, the
contact between scale and metal is lost at the edges and
corners. Hence the non-uniform oxidation occurred over the
surface. It was observed from the metallographic cross-
sections that the oxide layer was thicker and more uniform in
the middle than at the corners. Therefore, the oxide depth was
measured from the middle of the cross-section for each
sample. Figure 9 presents the oxide thickness from a series of
samples tested at 900
o
C but for different times of up to 695 hrs
for high-Si SiMo containing 4.8% Si, and 953 hrs for SiMo
containing 4.0% Si respectively. Neglecting the initial period,
it can be seen that the oxide penetration of both SiMo and
high-Si SiMo basically follows the linear growth in terms of
kinetics equations, which is typical for high-temperature
oxidation of metals. However, high-Si SiMo shows a thinner
oxide depth and smaller rate constant than SiMo. After 695
hours exposure the oxide depth of high-Si SiMo is 20 to 30%
less than that of SiMo samples.

The oxide nucleation and growth is illustrated as a function of
the exposure time in Figure 10, by taking the SiMo chemistry
as an example. After a short exposure time 2 hrs in Figure
10(a), the formed oxide consists of discrete, isolated particles
with a diameter of approximately 80 m. Each particle
contains multilayered scales, as indicated by different colors.
It appears that the oxide nuclei originated from the surface
defects which may be impurities, grain boundaries,
dislocations, or machining traces. Afterwards, the particles
grew laterally to yield a continuous film on the surface as
oxygen continued to dissolve and diffuse into the alloy phase,
as shown in Figure 10 (b) after 8 hours at 900
o
C. Further
increasing the testing time, continuous growth of the oxide
scales took place, as seen in Figures 10 (c), (d) and (e). The
micrographs show many voids and cracks in the scales. This
can be explained as follows. At high temperatures, the iron
ions and electrons migrated outwards and oxygen diffused
inwards to burn out graphite in the parent metal, thereby
resulting in some cavities. During cyclic oxidation testing,
thermal stress and strain were generated due to the different
thermal expansion and contraction between the scale and
metal (The coefficient of thermal expansion of iron oxides is
less than that of iron) and the volume ratio of oxide to metal.
This possibly made the oxides crack, collapse, and detach
from the metal matrix. Formation of porous zones and cracks
in turn alters the oxide adherence and oxidation kinetics. Note
that some cracks may be caused by sample preparation.
During the testing, no scale was spontaneously spalled off
from the samples. A stainless steel knife was used to
manually scrape the surface of oxidized specimens to examine
the oxide adherence. It was observed that many more oxide
debris were descaled from the SiMo samples, indicating that
the oxide layer of SiMo is not as adherent as that of high-Si
SiMo.

High-temperature oxidation of cast iron is more complex than
that of pure metals. The following is a brief discussion based
on our experimental results and theoretical descriptions [12,
13]. Cast irons are multi-component ferrous alloys, which
contain major (Fe, C, and Si), minor (< 0.1% such as S, Mg,
and Ce), and often alloying (> 0.1% such as Mo) elements.
Although some minor active elements may exert a significant
effect on the oxidation behavior [15], the influences of Fe, C,
Si, and Mo are examined in this paper.

Iron is the parent metal and forms with more than one oxide in
the scale. The oxide was found to consist of Fe
2
O
3

(haematite), Fe
3
O
4
, (magnetite), FeO (wustite), and
(FeO)
2
SiO
2
or Fe
2
SiO
4
(fayalite) when containing a certain
amount of Si [12, 13]. They have different structures and
formation mechanisms. The former two oxides were produced
during the outward oxidation while the fayalite, wustite, and
magnetite were formed during the inward oxidation, through a
series of inward/outward transport of oxygen ionizes, iron ions
and electrons. Due to the much greater mobility of defects in
FeO, this layer will be thicker than others [12]. The thin layer
Fe
2
O
3
at the outer surface was not discerned under optical
microscopy. In Figure 10, the two obvious layers should be
Fe
2
O
3
and FeO+Fe
2
SiO
4
respectively according to the relative
thickness and locations.

Regarding the effect of carbon, decarburization happened
through the reaction of graphite and oxygen, thus causing the
weight loss (negative weight change), as seen in Figure 8. We
conducted hot oxidation testing on SiMo irons with different
silicon contents ranging from 2.8 to 5.0%. The specimens
with higher Si contents seemingly had more weight loss at the
initial stage, while the weight loss was not observed for the
specimens with the Si content lower than 3.5%. This can be
explained as follows. As mentioned in the preceding
paragraph, the weight change determined in the testing
displayed the difference between the weight gain and loss.
For specimens with lower Si contents, the weight gain by
oxidation prevails over the weight loss by decarburization.
Therefore the total weight was monotonously increased for the
low-Si samples (< 3.5% Si) a function of exposure time. On
the other hand, less oxidation and thus more apparent weight
loss appeared for high-Si samples. Similar to other properties
of cast irons, oxidation resistance is significantly influenced
by the graphite structure. Oxidation and growth take place in
gray iron more rapidly than in compacted graphite iron (CGI)
and more rapidly in CGI than ductile iron [1].

Silicon is a major and economic element to improve high-
temperature oxidation resistance in cast iron. SiO
2
reacts with
FeO to form fayalite in close vicinity to the parent metal
surface. When the fayalite particles become larger they are
engulfed by the scale. These islands lie in the FeO layer as
markers and prevent the inward movement of the outer scale
[12]. The Si content can be increased to provide a denser
fayalite broken stringer on the metal substrate, and even a
continuous, protective SiO
2
scale. However, the mechanical
properties will be unacceptable when the Si content is too
high.

Oxide volatilization can occur in the Mo-O system at high
temperature and oxygen pressure [12]. Because of much lower
contents and oxygen affinities than other elements, and
formation of Mo-rich precipitates [9], Mo is much less likely
to be oxidized at 900
o
C than other elements in SiMo iron.

COEFFICIENT OF THERMAL EXPANSION - Dilatometry
testing was conducted to determine the critical temperature,
A1, and the coefficient of thermal expansion alpha . The
phase transformation of cast iron occurs in a temperature
range. Thus the critical temperature A1 can be defined as the
onset of the phase transformation from ferrite to austenite
upon heating, as indicated by inflections in the temperature-
alpha curves (Figure 11). It is known that A1 increases with
increasing heating rate because of the transformation kinetics.
In this work, the samples were slowly heated at 1 or 5
o
C/min.
There is no significant difference in alpha between 4.9% Si
SiMo and the regular SiMo with 4% Si when the temperatures
are below the A1. Clearly, the A1 temperature was raised by
approximately 60
o
C comparing the 4.9% Si with 4.0% Si
samples. This may imply that the practical operating
temperatures could be increased when replacing current SiMo
with high-Si SiMo according to the Al difference.

THERMAL FATIGUE RESISTANCE - In addition to high
temperature oxidation resistance and microstructure stability,
thermal fatigue life is also important for exhaust manifolds.
There are a few types of thermal fatigue tests including bars
[16], discs [17], and thin tubes [18] in terms of specimen
shape. Our Y-blocks were machined into bars and tested
under uniaxial conditions at CRS. The test method and
procedures were described in Reference 5. It is noteworthy
that the temperature range in the thermal fatigue testing was
from 500 to 950
o
C, instead of cooling to ambient temperature.
All data are listed in Table 1, which are fairly consistent for
each alloy. The Mo content was approximately 0.6% for all
samples. It is seen that the average number of cycles to failure
increased with increasing Si content from 4.0% to 4.5%, and
to 4.95%. It remains unclear why the high-Si SiMo displayed
a higher thermal fatigue life in this temperature range. From
the hot tensile data in Figure 5, silicon strengthening effect
almost vanished at high temperatures. However, in the light
of phase transformation, silicon raises A1 temperature and
widens the phase transformation range [1-4]. Therefore, the
effects of Si on the thermal fatigue resistance may be
attributed to the different amount of phase transformation
during the testing. Less amount of phase transformation
induced less thermal stress during the testing. It was reported
that some foundries produce manifolds and turbocharger
housings with spheroidal and compacted graphite irons at a
higher silicon content in order to improve the overall elevated-
temperature performance including oxidation, structure
stability, and thermal durability.

Table1. The number of cycles to failure of SiMo nodular
irons containing different amounts of Si in the uniaxial
thermal fatigue testing between 500
o
C and 950
o
C. The Mo
content was 0.6% in all samples.


Test # 4% Si 4.5% Si 4.95% Si
1 51 78 92
2 52 87 95
3 52 88 88
4 56 80 80
5 56 87
6 47 93
7 49
Average 52 83 90


PRELIMINARY PLANT TRIALS - Approximately 240
manifolds containing 4.5% to 5.0% Si were cast and machined
in the company plants to evaluate the rough handling and
machinability and identify the capability with the current
equipment. The SiMo production baseline was used,
including gating, risering, molding, melting, treatment, pour,
shake-out, cleaning, and machining parameters. It was
observed that there were rough handling issues when the Si
content is too high. For testing the durability and thermal
fatigue resistance of castings, Wescast Industries Inc. has built
a gas-fired engine exhaust simulator (EES). For comparison,
EES testing will be further conducted on SiMo and high-Si
SiMo irons for differently designed manifolds.

CONCLUSIONS
Alloy chemistry, solidification, microstructure and properties
of high-Si SiMo spheroidal cast irons have been investigated
in this work. Plant trials were performed to assess the
handling and machinability of high-Si SiMo manifolds.
Conventional Mg-Fe-Si treatment and inoculation can be
adopted for high-Si SiMo to produce satisfactory nodularity
and uniform nodule distribution. Casting experiments and
solidification modeling have revealed that the shrinkage
tendency of high-Si SiMo can be controlled to be the same
level as that of regular SiMo by using a suitable carbon
equivalent from 4.35 to 4.85 mainly depending upon the
critical thickness of castings.

Tensile UTS and YS are significantly increased with
increasing Si, while the strengthening effect of Si disappeared
at the testing temperatures higher than 500
o
C. Tensile
elongation at room temperature was decreased with increasing
Si content. However elongation of up 10% has been achieved
at room temperature when the silicon content approached 5%
by controlling the freezing behavior and nodule distribution.
Heat treatment can further increase tensile elongation. It was
observed that the impact toughness (unnotched charpy testing)
at room temperature of high-Si SiMo is much lower than that
of regular SiMo. Further work is needed to improve the
impact toughness of high-Si SiMo in order to alleviate
potential problems of casting handling prior to engine
assembly.

It has been repeatedly shown that high-Si SiMo is superior to
regular SiMo in high temperature oxidation and scaling
resistance tests. Uniaxial thermal fatigue life was increased
when silicon content was changed from 4.0 to 4.5 and to
4.95% when the testing temperatures were cycled between
500
o
C and 950
o
C.

ACKNOWLEDGMENTS
The results reported here would not have been possible
without the cooperative effort of the company leadership,
corporate and plants. Special thanks are due to B. Plank, G.
Burkhalter, J. Cassidy, A. Cormier, G. Liao, B. Kowal, F. Yu,
P. Slater, D. Lanting, P. Pickard and A. Valentyn for technical
support, and T. Thoma and C. Sloss for their helpful review
and comments. The authors also thank P. Chan and R.
Gundlach of CRS for the thermal fatigue testing and
discussions.

REFERENCES
1. J.R. Davis (editor), ASM Specialty Handbook: Cast
Irons, ASM International, Materials Park, OH, 1996.
2. G.M. Goodrich (technical editor), Iron Castings
Engineering Handbook, AFS (American Foundry
Society), Des Plaines, IL, 2003.
3. R. Elliott, Cast Iron Technology, Butterworths, London,
UK, 1988.
4. H.T. Angus, Cast Iron: Physical and Engineering
Properties, Butterworths, London, UK, 1976.
5. R.B. Gundlach et al., Thermal Fatigue Resistance of
Silicon-Molybdenum Ductile Irons, Presentation at
DIS/AFS Millis Symposium, October 1998, Hilton Head,
SC.
6. D.L. Sponseller, W.G. Scholz, and D.F. Rundle,
Development of Low-Alloy Ductile Irons for Service at
1200-1500 F AFS Trans., 76 (1968), 353-368.
7. K. Kampendonk, R. Williams, and R. Perrin, Private
Communication, Wescast Industries Inc. July, 2003.
8. M.D. Chaudhari, R.W. Heine, and C.R. Loper,
Principles Involved in the Use of Cooling Curves in
Ductile Iron Process Control, AFS Trans., 82 (1974),
431-440.
9. B. Black et al., Microstructure and Dimensional
Stability in Si-Mo Ductile Irons for Elevated
Temperature Applications, Proc. of the 2002 SAE Inter.
Body Eng. Conf. and Automotive & Transportation
Tech. Conf., Paris, France (paper # 2002-01-2115).
10. C.A. Bhaskaran and D.J. Wirth, Ductile Iron Shrinkage
Evaluation through Thermal Analysis, AFS Trans., 110
(2002), 1-16 (paper # 02-003).
11. Foundry Handbooks, Vol. 1: Cast Irons, Mechanical
Engineering Publisher, Beijing, 2002 (in Chinese).
12. N. Birks and G.H. Meier, Introduction to High
Temperature Oxidation of Metals, Edward Arnold
(Publishers) Ltd, London, UK, 1983.
13. P. Kofstad, High Temperature Corrosion, Elsevier
Applied Science Publisher Ltd., Essex, England, 1988.
14. E.A. Loria, Cyclic Oxidation of Chromized Steel and
Competitive Materials at 650 to 815
o
C, J. Mater. for
Energy Systems, 8 (1986) 132-141.
15. E. Lang (editor), The Role of Active Elements in the
Oxidation Behavior of High Temperature Metals and
Alloys, Elsevier Applied Science, Essex, England, 1989.
16. K. Akiyama et al., Analysis of Thermal Fatigue
Resistance of Engine Exhaust Parts, SAE Trans.
910430, 1991, 63-71
17. A. Weronski, Thermal Fatigue of Metals, Marcel Dekker
Inc, New York, NY, 1974.
18. M.H. Aliabadi (editor), Thermomechanical Fatigue and
Facture, WIT press, Boston, MA, 2002.









2.9
3.0
3.1
3.2
3.3
3.4
4.2 4.4 4.6 4.8 5.0 5.2
Trial Data
C+Si/3 = 4.6
C+Si/3 = 4.7
C+Si/3 = 4.8
Uniform
Nodule
Primary
Nodule
Carbon Flotation
and Shrinkage
Shrinkage
and Carbides
Silicon Content %
C
a
r
b
o
n

C
o
n
t
e
n
t

%


Figure 1: Silicon, carbon, and the carbon equivalent (CE) map for high-Si SiMo. The circles stand
for some trial chemistries. Prediction of microstructure and shrinkage is presented.


1050
1090
1130
1170
1210
1250
0 60 120 180 240 300 360
Time (s)
T
e
m
p
e
r
a
t
u
r
e

(
C
)















Figure 2: Measured cooling curves and the resulting microstructures. Curve (a) corresponds to
micrograph (b), and curve (c) to micrograph (d).
(a)
(c)
(b) (d)
200 m 200 m
























Figure 3: Casting experiments and modeling of shrinkage for high-Si SiMo of 4.5% Si. The C content is 3.35%, i.e.
CE = 4.85 for the top row of pictures, while the C content is 3.15%, i.e. CE = 4.65 for the bottom pictures. The
pictures in the left, middle, and right columns represent the flange sections of manifolds, sections of AFS blind
risers, and solidification modeling respectively.


































400
500
600
700
800
3.50 4.00 4.50 5.00 5.50
Si (wt.%)
S
t
r
e
n
g
t
h

(
M
P
a
)
YS
UTS
(a) 4
8
12
16
20
3.50 4.00 4.50 5.00 5.50
Si (wt.%)
E

(
%
)

(b)
150
180
210
240
270
300
3.50 4.00 4.50 5.00 5.50
Si (wt.%)
H
B
W
(c)
Figure 4: Silicon content versus room temperature
properties: (a) tensile UTS and 0.2% offset YS,
(b) tensile elongation E%, and (c) Brinell
hardness HBW. Mo content varied from 0.6% to
0.9%. Each point represents three tests at least.



0
150
300
450
600
3.5 4.0 4.5 5.0 5.5
Si (wt.%)
U
T
S

(
M
P
a
)
450C 700C 900C



400
500
600
700
800
as-cast subcritical full
Sample Condition
S
t
r
e
n
g
t
h

(
M
P
a
)
13
14
15
16
17
E

(
%
)
Yield UTS E%



0
5
10
15
20
25
0.5 2 3.5
E
n
e
r
g
y

A
b
s
o
r
b
e
d

(
J
)
4.65% Si 5.01% Si
As-cast Sub-critical Full Annealing




Figure 5: Silicon content versus the hot
tensile UTS tested at different
temperatures. Each point represents the
average of three tests at least. The points
are connected just for showing the
tendency to change. Mo content is 0.75%.


Figure 6: Room temperature tensile
results of as-cast, subcritically
annealed, and full annealed
specimens of high-Si SiMo (4.65%
Si and 0.75% Mo).

Figure 7: Absorbed energy of non-
notched charpy testing for 4.65% Si
and 5.01% Si SiMo samples with
different treatment: as-cast,
subcritical annealing, and full
annealing respectively. The sample
dimension is 1010 55 mm.



-5
0
5
10
15
20
25
0 200 400 600 800
Time at Temperature of 900 C (hr.)
W
t
.

C
h
a
n
g
e

(
m
g
/
c
m
^
2
)



Figure 8: Weight change rate versus time obtained from hot oxidation testing at 900
o
C. Squares and
circles represent SiMo of 4% Si and high-Si SiMo of 4.8% Si respectively. Solid and empty symbols
stand for airflow of 2.5 liters/min through the furnace and no airflow introduced respectively.


0
200
400
600
800
1000
1200
0 200 400 600 800 1000
4.0% Si
4.8% Si
Time at 900
o
C (hr)
O
x
i
d
e

D
e
p
t
h

(

m
)



Figure 9: Oxide depth measured from the cross-section of tested samples using optical
microscopy. The vertical bars represent the measurement variations of the oxide thickness. The
points are connected for showing the tendency to change.


































Figure 10: Cross-section micrographs of SiMo samples with 4% Si after cyclic oxidation testing at 900
o
C for
different exposure times (hours) in total: (a) 2, (b) 8, (c) 24, (d) 415, and (e) 695. Airflow of 2.5 liters/min was
introduced through the furnace. Same scale was used for the five pictures. For each micrograph, the far left and
right edges represent the mounting material and the iron matrix respectively, and the middle regions represent the
oxide layers.



-20
-10
0
10
20
800 840 880 920 960 1000
Temperature (C)
A
l
p
h
a
,

1
0
^
(
-
6
)



200 m
(b)
(c)
(d)
(e)
(a)
Fe
3
O
4

FeO+Fe
2
SiO
4

Figure 11: Coefficient of thermal
expansion alpha versus temperature
measured from dilatometer testing at a
heating rate of 5
o
C/min: (a) SiMo of
4% Si and (b) high-Si SiMo of 4.9%
Si. The critical temperature A1 can be
determined from the curves.

(a)
(b)

Você também pode gostar