You are on page 1of 9

Effect of Prior-Austenite Grain Size and Transformation

Temperature on Nodule Size of Microalloyed


Hypereutectoid Steels
A.M. ELWAZRI, P. WANJARA, and S. YUE

The effect of prior-austenite grain size and transformation temperature on nodule size and colony size
of hypereutectoid steels containing 1 pct carbon with different levels of vanadium and silicon was
investigated. Specimens of the various steels were thermally processed at various temperatures ranging
from 900 °C to 1200 °C and transferred to salt bath conditions at 550 °C, 580 °C, and 620 °C to
examine the structural evolution of pearlite. The heat-treatment work showed that for only the
hypereutectoid steel without vanadium there was a continuous grain boundary cementite network,
the thickness of which increased with increasing reheat temperature. Analysis of the thermally processed
hypereutectoid steels also indicated that the prior-austenite grain size and transformation temperature
controlled the nodule size, while the colony size was dependent on the latter only.

I. INTRODUCTION process. This plausible progression leads to the transforma-


tion of austenite to a lamellar product that has a specific spa-
THERE has been a significant interest in the addition of tial orientation in which the ferrite and cementite phases are
microalloying elements, particularly vanadium, to medium and parallel to each other within a colony region. In reality, for
high carbon steels with predominantly pearlitic microstruc- each colony of pearlite, the lamellae are mostly parallel and
tures,[1,2,3] since it has been recognized that the formation of a are frequently curved. However, different pearlite colonies
fine dispersion of vanadium carbide precipitates within the have different lamellae orientations, and, as the transforma-
pearlitic ferrite lamellae increases strengthening.[1,4,5] Similarly, tion progresses, neighboring colonies of lamellae join together
Tarui et al.[6] have shown that increasing the silicon content and continue to advance into the austenite, such that when
from 0.2 to 1.5 pct in an eutectoid carbon steel improved both the transformation occurs at low degrees of undercooling the
the tensile yield and ultimate strengths, primarily due to solid colony groups advance by a boundary that is roughly spher-
solution strengthening of the pearlitic ferrite by silicon addi- ical in shape, leading to the formation of a pearlite nodule.
tion. In these cases, the increased strengthening in microalloyed Eventually, as the nodules continue to grow, they impinge on
steels has been related to the main microstructural constituent, one another to complete the transformation. If the temperature
pearlite, which consists of two phases, ferrite and cementite, as at which the transformation is occurring is decreased, such
alternating lamellae. Generally, pearlite is considered to be a that the undercooling enables the nucleation process to become
transformation product of the decomposition of austenite by a so rapid that all of the grain boundaries are consumed very
diffusional process involving nucleation followed by growth.[7–11] early in the transformation, the pearlite nodules are then
The initiation of a pearlitic microstructure from austenite observed to be in contact along the prior-austenite boundaries
entails the formation of neighboring nuclei of ferrite and as a continuous cementite network. In previous work, Han
cementite on an austenite grain boundary. Supposing that ini- et al.[12,13] and Khalid and Edmonds[14] have shown that
tially only one nucleus of cementite forms on this boundary, vanadium additions can prevent the formation of a continu-
this region will be locally rich in carbon, which has obtained ous cementite network along prior-austenite grain boundaries
from the immediate surroundings. This then reduces the car- in hypereutectoid steels with carbon concentrations above
bon content on either side of the cementite nucleus and encour- 0.8 wt pct. However, the influence of materials and process-
ages the adjacent nucleation of ferrite, as shown in Figure 1. ing conditions, particularly the role of microalloying additions
If this occurs and the cooperative process continues, adja- and reheat temperature, on the grain boundary cementite net-
cent nuclei of alternating ferrite and cementite are formed. work and pearlite colony characteristics were not addressed
These can then grow by a relatively short-range diffusion in their work. The present study was thus undertaken to exam-
process in which carbon diffuses parallel to the reaction front, ine the influence of thermal processing conditions (reheat and
from in front of the growing ferrite to the growing cemen- transformation temperatures) on the grain boundary cemen-
tite. More importantly, the diffusion distance does not increase tite network and nodule size of the pearlitic structure in hyper-
with time, as would be the case for a continuous precipitation eutectoid steels with different levels of vanadium and silicon.

A.M. ELWAZRI, Research Associate, and S. YUE, Professor, are with II. EXPERIMENTAL PROCEDURE
the Department of Metals and Materials Engineering, McGill University,
Montreal, PQ, Canada H3A 2B2. Contact e-mail: abdelbaset.el-wazri@mcgill.ca A. Material
P. WANJARA, Research Officer, is with the Aerospace Manufacturing
Technology Centre, Institute for Aerospace Research, National Research
Three hypereutectoid steels (A, B, and C) with varying sil-
Council of Canada, Montreal, PQ, Canada H3T 2B2. icon, vanadium, and nitrogen contents were prepared at CAN-
Manuscript submitted May 14, 2004. MET Materials Technology Laboratory (Ottawa, ON, Canada),

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, SEPTEMBER 2005—2297


In this way, a range of different austenite grain sizes was
obtained. To minimize oxidation during the heat treatment,
the specimen was enclosed in a quartz tube sealed with
O-rings in which a high-purity argon atmosphere was main-
tained. Quenching the sample was done by a lever, through
a hole, that was made in the lower support between the inside
wall of the quartz tube and the lower anvil. The opening was
(a) (b) kept closed throughout the experiment and was opened for
quenching of the specimen after austenitization. This mech-
anism permitted quenching of the specimens within 1 second.
In particular, the specimens were quenched into a salt bath
for transformation of the austenite phase at temperatures of
550 °C, 580 °C, and 620 °C. Complete transformation of
austenite to pearlite was observed to occur within 10 minutes
of placing the specimens in the salt bath.

(c) (d) D. Metallography


Fig. 1—Schematic representation of pearlite formation by nucleation and Metallographic preparation of the steel specimens involved
growth: (a) through (d) successive steps in time sequence.
mounting in bakelite followed by automated grinding using
successively finer silicon carbide papers from 60 to 800 grit.
The specimens were then polished at a platen speed of
Table I. Chemical Composition of Experimental Steels
120 rpm with 9, 3, and 1 m water-based diamond sus-
(Weight Percent)
pensions on short nap cloths with an alcohol-based lubricant.
Steels A B C D To reveal the prior-austenite grain boundaries, the austeni-
C 1.1 1.0 1.08 0.91 tized and quenched specimens were etched for 30 seconds
Si 0.23 0.99 0.78 0.22 by immersing in a solution containing 4 drops HCl, 10 drops
Mn 0.63 0.77 0.75 0.49 wetting agent Teepol (BDH Chemical Limited, Poole,
Cr 0.04 0.066 0.049 0.04 England) and 100 mL saturated aqueous picric acid that was
Ni 0.038 0.039 0.039 0.07 heated to 80 °C. During etching, the specimens were swabbed
Cu 0.036 0.037 0.037 0.02 regularly with cotton wool to remove dark deposits from
V 0.17 0.078 0.26 — forming due to the chemical attack. The prior-austenite grain
P 0.010 0.0086 0.009 0.006 size of the specimens was then determined from light optical
S 0.007 0.008 0.009 0.005 micrographs by the intercept method.[15]
N 0.007 0.013 — 0.0031
Alternatively, the general microstructures of the steels
austenitized at different temperatures were revealed by
immersing in a 3 pct nital etchant and examined using a
using vacuum melting. A fourth hypereutectoid steel without light optical microscope and a field-emission gun scanning
vanadium (D), used as a reference, was provided by IVACO electron microscope (FEGSEM). The delineated pearlite nod-
Rolling Mills (L’Orignal, ON, Canada). The chemical com- ules (Figure 2) were then measured from a series of light
positions of these steels are given in Table I. From each steel, optical micrographs using the linear intercept method in
cylindrical specimens were machined to have dimensions of which the mean linear intercept, l , was the average size of
11.4 mm in height and a 1.5 height-to-diameter ratio. the nodules and computed as[16]
l
B. Continuous Cooling Compression Testing l [1]
MN
To characterize the austenite-to-pearlite transformation,
continuous cooling compression (CCC) testing was performed where l is the total line length employed, M is the magni-
by heating the steel specimens to the austenitizing temperature, fication of the micrograph, and N is the number of nodule
and holding for 20 minutes for thermal stabilization, followed boundaries intersected by the test line.
by cooling at a constant rate of 1 °C  s1 to 850 °C. At this
temperature, continuous deformation at a constant strain rate
of 0.001 s1 was applied to the specimen while maintaining III. RESULTS AND DISCUSSION
cooling at a rate of 1 °C  s1. The true stress and true strain A. CCC Analysis
values were calculated from the load-displacement data gen-
erated during deformation. The CCC method is based on the correlation of significant
changes in the flow stress vs strain behavior to microstructural
changes. Figure 3 shows the hypothetical flow stress behavior
C. Heat Treatments as a function of temperature during CCC testing conditions
Heat treatment of the specimens was performed by heating where there is no microstructural change. Therefore, under
at a constant rate of 1.5 °C  s1 to various temperatures of practical conditions, any deviation from a smooth increase
900 °C, 1000 °C, 1100 °C, and 1200 °C, followed by holding in the flow stress with decreasing temperature can be related
for 20 minutes to permit full austenitization of the material. to a change in the microstructure of the material.

2298—VOLUME 36A, SEPTEMBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


(a)

Fig. 3—Hypothetical flow stress behavior during CCC testing.

(b)
Fig. 2—Typical microstructure of a microalloyed hypereutectoid steel show-
ing (a) the pearlite colonies and nodules, which have been delineated in
(b) using a thick white line to define the pearlite nodule boundary and a
thin line to define a pearlite colony boundary within the nodules.
Fig. 4—True stress-strain curves of steels A through D obtained by CCC
testing. Arrows labeled 1 and 2 represent the first and second deviations.

For the steels A, B, C, and D, the CCC true stress-strain


curves, given in Figure 4, were observed to exhibit an over- as the latter is softer than the parent phase at any given tem-
all increase in the flow stress with increasing strain and perature.[18] As shown in Figure 4, a second deviation, which
decreasing temperature with specific deviations occurring at occurred at a strain of 0.3  0.05, was characterized by a
different temperatures and flow stress levels that reflected change back to a positive slope in the stress-strain curve,
the compositional differences in these microalloyed steels. inevitably due to the completion of the austenite transfor-
In particular, the addition of vanadium and silicon (steels mation (Ar1) that then produces a rapid increase in the
A, B, and C, vs D) was determined to significantly increase strengthening rate after the first deviation.
the flow stress levels in these steels, especially at tempera- From the CCC results, the transformation start tempera-
tures below 800 °C. This flow stress increase may be related tures (Ar3) were obtained to be about 620 °C, 630 °C, 645 °C,
to the solute drag and solid solution strengthening effects of and 600 °C for steels A, B, C, and D, respectively. Steel C
vanadium and silicon that have been noted to retard dynamic was determined to have a higher Ar3 (i.e., faster austenite to
recovery.[17,18] Additionally, for each flow stress-strain curve, pearlite transformation kinetics) as compared to steels A, B,
two deviations or inflection points were observed, which and D, most likely due to the highest vanadium content and
could be related to the Ar3 and Ar1 temperatures. The first relatively high silicon level (0.8 pct) in the former mater-
deviation in Figure 4, characterized by a stress drop and neg- ial. Hence, from the CCC curves, it is evident that silicon
ative slope in the flow stress-strain curve, indicated that the and vanadium additions, either separately or in combination,
material became softer and corresponds to the start temper- can have a noticeable effect on the transformation tempera-
ature for the transformation of austenite to pearlite (Ar3), ture in microalloyed steels. Based on these transformation

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, SEPTEMBER 2005—2299


Fig. 5—Reheated austenitic microstructures of steel D.

temperature findings, the salt bath conditions of 550 °C,


580 °C, and 620 °C were selected to investigate the evolution
of the pearlitic microstructure with thermal processing.

B. Reheated Austenite Grain Size


The prior-austenite grain structure, exemplified in Figure 5
for steel D, was analyzed for grain size quantification, the
results of which are given in Figure 6 for the various mate-
rials and reheat temperature conditions. As expected, the
grain size increased with increasing austenitizing tempera-
ture from 900 °C to 1200 °C. Also, the inhibitory effect of
vanadium and silicon on the growth of prior-austenite grains
can be readily observed by noting the progressively smaller
grain size values for steels A, B, and C as compared to the
larger values for steel D, which had no microalloying addi-
Fig. 6—Effect of austenitizing temperature on prior-austenite grain size.
tions. This refinement in the prior-austenite grain structure
can be attributed to the presence of incompletely dissolved
vanadium carbide particulates in the ferritic lamellae[20,21,22]
of the pearlite structure, which enabled effective pinning of formation. This is exemplified through the austenitizing con-
the grain boundary, or it could be due to solute drag effects. dition at the lower reheat temperature of 900 °C (Figures 7(a)
through (c)), where the prior-austenite grain size was 48 m,
C. Heat-Treatment Effect on Cementite Network which resulted in a thinner layer of grain boundary cemen-
tite as compared to reheating to 1200 °C, as shown in
1. Hypereutectoid steel without vanadium and silicon Figure 7(d), where the grain size was more than double
From the microstructures of the heat-treated vanadium- (110 m).
free steel D at the different temperatures, the evolution of
the prior-austenite grain structure characteristics was exam- 2. Hypereutectoid steel with vanadium and silicon
ined for the influence of thermal processing conditions, as The addition of microalloying elements to a hypereutectoid
illustrated in optical micrographs given in Figure 7. Increas- steel with a low or high silicon content was observed to
ing the reheat temperature was observed to increase not only prevent the formation of a thick continuous cementite layer
the prior-austenite grain size, but also the thickness of the at the prior-austenite grain boundaries, as illustrated in the
grain boundary cementite network that was observed to form optical micrographs given in Figure 8 for steel A. The nature
in steel D for all the heat treatments examined in this work. of the grain boundary cementite for the microalloyed steels
As given in Table II, there is a progressive increase in the was difficult to examine even from the general microstruc-
thickness of the cementite network with increasing reheat tural observations performed using an FEGSEM at relatively
and transformation temperatures, and this may be related to low magnifications (3000 times), as shown in Figure 9. In
the increased kinetics for grain boundary cementite forma- fact, low voltage imaging was employed on a FEGSEM at
tion at higher temperatures. Moreover, with increasing reheat high magnifications (20,000 times) to ascertain the presence
temperature, there is an increase in the prior-austenite grain of the grain boundary cementite particles, as depicted in
size, as given in Figure 6, which will increase the thickness Figure 10.
of the cementite network through a reduction in the overall For hypereutectoid steels, the occurrence of a continuous
grain boundary area available for proeutectoid cementite grain boundary cementite network renders considerable

2300—VOLUME 36A, SEPTEMBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


(a) (b)

(c) (d)
Fig. 7—Optical micrographs showing the microstructure of steel D: (a) austenitized at 900 °C, salt bath at 550 °C; (b) austenitized at 900 °C, salt bath at
580 °C; (c) austenitized at 900 °C, salt bath at 620 °C; and (d) austenitized at 1200 °C, salt bath at 620 °C.

Table II. Thickness of the Continuous Cementite Network presence of vanadium carbides, in small quantities, at the
in Steel D prior-austenite grain boundaries, with the majority of these
being precipitated in the grain boundary ferrite. This occur-
Heat-Treated
Heat Transformation Thickness of rence of vanadium carbide precipitation has then been sug-
Treatment (°C) Temperature (°C) Cementite Network gested by Han et al.[24] to have a twofold effect on grain
boundary cementite formation. An effect of the vanadium
1200 620 25.5  3 m
carbide formation on the austenite grain boundary is the local
580 20.5  2 m
550 16.8  2 m depletion of carbon that reduces its diffusion along the bound-
900 620 20.0  3 m ary to the cementite particles, thereby inhibiting their rapid
580 19.9  3 m growth into the boundary.[14] An additional effect of the car-
550 14.0  2 m bon being coupled into precipitation is a concomitant pro-
motion of ferrite nucleation and growth, which aids further
in inhibiting the continuance of grain boundary cementite.
The findings of this current work also suggest an influence
difficulty for cold forming operations, such as drawing, since of silicon on the effectiveness of vanadium microalloy addi-
the brittle cementite layer provides an easy crack path for tions to refine structural characteristics in hypereutectoid steels.
fracture.[3,12–14,17] The ability to restrict the grain boundary In particular, it was observed that a high content of silicon
cementite network from linking through vanadium micro- (0.99 pct) provided a slight decrease in the prior-austenite
alloying has been attributed to an increase in the driving grain size even at a low vanadium level (0.078 pct), as illus-
force for carbide nucleation in the presence of vanadium, trated by the finer grain size of steel B as compared to A in
which leads to additional carbide nuclei at austenite grain Figure 6. In the work of Han[25] and Wada et al.,[26] the addi-
boundaries.[5,23] Moreover, Han et al.[24] have observed the tion of silicon was determined to decrease the solubility of

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, SEPTEMBER 2005—2301


(a) (b)

(c) (d)
Fig. 8—Optical micrographs showing the microstructure of steel A: (a) austenitized at 1200 °C, salt bath at 620 °C; (b) austenitized at 1200 °C, salt bath
at 550 °C; (c) austenitized at 900 °C, salt bath at 620 °C; and (d) austenitized at 900 °C, salt bath at 550 °C.

Fig. 9—FESEM images showing the microstructure of steels A and C austenitized at 1200 °C and transformed at 620 °C.

vanadium carbide particles in both austenite and ferrite by austenite and ferrite phases.[13] In addition, the lower vanadium
raising the carbon activity. This suggests that in the presence carbide solubility in the presence of silicon suggests that
of silicon, vanadium carbide formation is promoted in the precipitation can occur at higher austenitizing temperatures.

2302—VOLUME 36A, SEPTEMBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 10—Low-voltage imaging using an FESEM at higher magnifications to show the grain boundary cementite particles in steel A austenitized at 1200 °C
and transformed at 620 °C.

Fig. 11—Effect of reheat temperature on the nodule size for the hypereu- Fig. 12—Effect of prior-austenite grain size on the nodule size for the
tectoid steels A through D processed at a transformation temperature of hypereutectoid steels A through D processed at a transformation temperature
620 °C. of 620 °C.

This would promote austenite grain size refinement as observed materials and processing parameters on the prior-austenite
in the present work with the vanadium microalloyed steels grain size. In particular, the austenitizing temperature deter-
having different silicon contents. mines the size of the prior-austenite grains, whose bound-
aries in turn provide nucleation sites for the pearlite nodules.
Hull et al.[8,11] have reported that the nucleation rate of pearlite,
D. Heat-Treatment Effect on Nodule Size hence nodule diameter, is very sensitive to the variation in
To examine the influence of thermal processing on pearlite prior-austenite grain size. Therefore, for a specific transfor-
nodule characteristics, the size of the nodules was plotted as mation temperature, which dictates the number of nucleation
a function of the reheat temperature, as shown in Figure 11 sites for pearlite formation, the nodule size should be directly
for the various hypereutectoid steels (A, B, C, and D) at a related to prior-austenite grain size, the interdependence of
specific transformation temperature of 620 °C. By comparing which is plotted in Figure 12.
the behaviors of steels A and D, it can be observed that, at It is then not surprising that in the plots of nodule size as
a given transformation condition, the addition of vanadium a function of the transformation temperature, as illustrated
enables a reduction in the nodule size for a particular reheat in Figures 13 and 14 for steels D and A, respectively, four
temperature. Further reductions in the pearlite nodule size independent curves were assembled according to prior-
were observed with increasing silicon and vanadium contents, austenite grain size. Between 550 °C and 620 °C, a linear
as shown in Figure 11 for steels B and C, respectively. The relationship between nodule size and transformation tem-
progressive decrease observed for the nodule size with perature was observed for all the hypereutectoid steels
decreasing reheat temperature and increasing vanadium or sil- reheated to temperatures between 900 °C to 1200 °C.
icon additions may be inherently related to the effects of these This dependence of the nodule size on the transformation

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, SEPTEMBER 2005—2303


Fig. 13—Effect of the transformation temperature on nodule size for dif-
ferent prior-austenite grain sizes (reheat temperature) for steel D.
Fig. 15—Effect of the prior-austenite grain size on average pearlite colony
size in steels A through D at a transformation temperature of 620 °C.

Fig. 14—Effect of transformation temperature on nodule size for different


prior-austenite grain sizes (reheat temperature) for steel A.

Fig. 16—Effect of transformation temperature on average pearlite colony


conditions is most likely linked to the increase in the prior- size in steels A through D austenitized at 1200 °C.
austenite grain size with increasing transformation temper-
ature, as shown in Figures 7 and 8, since Marder and
Bramfitt[27,28] have shown through cinephotomicrography ture, the pearlite was shown to have a clear nodular
using a hot-stage nodules that nucleate only at austenite grain microstructure, while with a finer austenite grain size, the
boundaries. pearlite was comprised of individual “units” limited to a few
colonies.[30] Hence, Garbarz and Pickering[29] have reported
that the austenite grain size has an effect on the transfor-
E. Heat-Treatment Effect on Pearlite Colony Size mation of the pearlitic structure from a nodular appearance
The effect of reheat temperature on the pearlite colony at a grain size of 180 m to one that is colony-like for grain
structure was investigated by plotting the pearlite colony sizes less than 25 m. In the present work, for the isother-
size as a function of the prior-austenite grain size for steels mal temperature condition of 900 °C, it was observed that
A, B, C, and D, as shown in Figure 15 for the transformation the nodular structure was slightly more difficult to identify
temperature of 620 °C. Since the observed change in pearlite for the microalloyed steels, probably due to the occurrence
colony size with prior-austenite grain size was within the of this transformation to a colony base structure because of
0.6-m error in size measurement, a negligible effect was the finer austenite grain size (near 30 m).
concluded for all the hypereutectoid steels investigated. This In the present work, the effect of the transformation tem-
finding is in agreement with previous work by Pickering and perature on the pearlite colony size was determined for steels
Garbarz,[3] who have reported that the prior-austenite grain A, B, C, and D, as illustrated in Figure 16 for austenitizing
size affects the morphology of pearlite nodules rather than conditions of 1200 °C. For all the hypereutectoid steels
the size of the colonies. Specifically, Garbarz and Pickering[29] examined, the pearlite colony size increased with increas-
have shown that the dependence of the pearlite colony size ing transformation temperature and the relationship appears
on the transformation temperature is similar to that of the to be similar to that reported in the work of Garbarz and
interlamellar spacing variation with the transformation con- Pickering.[29] The variation in the pearlite colony size with
ditions. In this work, for a coarse prior-austenite grain struc- the transformation temperature has been attributed to the

2304—VOLUME 36A, SEPTEMBER 2005 METALLURGICAL AND MATERIALS TRANSACTIONS A


dependence of the interlamellar spacing on cooling rate from REFERENCES
the reheat temperature and the isothermal transformation 1. F.A. Khalid, D.A. Gilroy, and D.V. Edmonds: Processing, Microstructure
temperature, of which only the latter was relevant for the and Properties of Microalloyed and Other Modern High Strength Low
present work. In particular, Marder and Bramfitt[27,28] have Alloy Steels, Proc., Pittsburgh, PA, 1992, A.J. DeArdo, ed., TMS-AIME,
indicated that the apparent mechanism for pearlite transforma- Warrendale, PA, 1992, pp. 67-75.
2. E.E. Laufer and D.M. Fegredo: Can. Met. Q., 1983, vol. 22, pp. 193-99.
tion is volume diffusion, and thus the interlamellar spacing, 3. F.B. Pickering and B. Garbarz: Mater. Sci. Technol., 1989, vol. 5,
along with the pearlite colony size, is dependent on the pp. 227-37.
growth velocity, which in turn is related to the transforma- 4. G.L. Dunlop, C.J. Carlsson, and G. Frimodig: Metall. Trans. A, 1978,
tion temperature. vol. 9A, pp. 261-66.
5. T.D. Mottishaw and G.D.W. Smith: HSLA Steels Technology and Appli-
cations, Inst. Conf. Proc., ASM INTERNATIONAL, Philadelphia, PA,
1984, pp. 163-75.
IV. CONCLUSIONS 6. T. Tarui, T. Takahashi, S. Ohashi, and R. Uemori: I&SM, 1994, pp. 25-31.
7. J.R. Vilella, G.E. Guellich, and E.C. Bain: 17th Annual Convention of
1. The Ar3 was found to increase with increasing vanadium the Society, Transactions of the A.S.M., Chicago, IL, 1935, pp. 225-61.
and silicon content. 8. F.C. Hull and R.F. Mehl: Trans. ASM, 1942, vol. 30, pp. 381-424.
2. For the hypereutectoid steels with microalloy additions 9. N.T. Belaiew: J. Iron Steel Inst., 1922, vol. 60, pp. 201-39.
(A, B, and C), the transformation temperatures were 10. G.E. Pellisier, M.F. Hawkes, W.A. Johnson, and R.F. Mehl: Trans.
determined to be about 30 °C higher than in a similar steel ASM, 1942, vol. 29, pp. 1049-86
11. F.C. Hull, R.A. Colten, and R.F. Mehl: Trans. AIME, 1942, vol. 150,
grade without vanadium (D). pp. 185-207.
3. It was observed that the addition of vanadium significantly 12. K. Han, T.D. Mottishaw, G.D.W. Smith, D.V. Edmonds, and A.G.
increased the flow stress of the hypereutectoid steels, Stacey: Mater. Sci. Eng, 1995, vol. 190A, pp. 207-14.
particularly at temperatures below 800 °C. 13. K. Han, T.D. Mottishaw, G.D.W. Smith, and D.V. Edmonds: Mater.
Sci. Technol. 1994, vol. 10, pp. 955-63.
4. In hypereutectoid steel without vanadium (D), the prior- 14. F.A. Khalid and D.V. Edmonds: Scripta Metall. Mater., 1994, vol. 30,
austenite grain boundaries consisted of a continuous pp. 1251-55.
cementite network, the thickness of which was dependent 15. ASTM standard E112-96, “Standard Test Methods for Determining
on the reheat and transformation temperatures. Average Grain Size,” Annual Book of ASTM Standards, ASTM,
5. For the microalloyed hypereutectoid steels (A, B, and C), Philadelphia, PA, 1996, vol. 03.01, pp. 251-74.
16. E.E. Underwood: Quantitative Stereology, Addison-Wesley, Reading,
a discontinuous film of cementite was apparent at the MA, 1970, pp. 73-75.
prior-austenite grain boundary, most likely due to the role 17. R.W.K. Honeycombe: Steels Microstructure and Properties, Edward
of vanadium and silicon on severing the cementite network Arnold, London, 1981.
through vanadium carbide precipitation. 18. S. Serajzadeh and A.K. Taheri: Mater. Lett., 2002, vol. 56, pp. 984-89.
19. S. Zajac, T. Siweck, W.B. Hutchinson, and R. Lagneborg: Iron Steel
6. The pearlite nodule size was observed to decrease with Inst. Jpn. Int., 1998, vol. 38, pp. 1130-39.
increasing vanadium and silicon additions and increase 20. A.M. Elwazri, P. Wanjara, and S. Yue: Mater. Sci. Eng., 2003,
with both the transformation and reheat temperatures, vol. 339A, pp. 209-15.
inherently due to the dependence of the prior-austenite 21. A.M. Elwazri, P. Wanjara, and S. Yue: Iron Steel Inst. Jpn., 2004,
grain size on these materials and processing variables. vol. 45, pp. 162-70.
22. A.M. Elwazri: Ph.D. Thesis, McGill University, Montreal, 2004.
7. The pearlite colony size was determined to be dependent 23. J.G. Zimmerman, R.H. Aboron, and E.C. Bain: Trans. ASM, 1937,
on the transformation temperature and independent of the vol. 25, pp. 755-87.
prior-austenite grain size. 24. K. Han, G.D.W. Smith, and D.V. Edmonds: Metall. Mater. Trans. A,
1995, vol. 26A, pp. 1617-31.
25. K. Han: Scripta Metall., 1993, vol. 28, pp. 699-702.
ACKNOWLEDGMENTS 26. T. Wada, H. Wada, J.F. Elliot, and J. Chipman: Metall. Trans., 1972,
vol. 3, pp. 1657-62.
The authors thank the Canadian Steel Industry Research 27. A.R. Marder and B.L. Bramfitt: Metall Trans., 1975, vol. 6, pp. 2009-14.
28. A.R. Marder and B.L. Bramfitt: Metall Trans., 1976, vol. 7, pp. 365-72.
Association (CSIRA) and the Natural Sciences and Engine- 29. B. Garbarz and F.B. Pickering: Scripta Metall., 1987, vol. 21, pp. 249-53.
ering Research Council of Canada (NSERC) for their finan- 30. B. Garbarz and R. Kuziak: Microalloying ’95, ISS, Pittsburgh, PA,
cial support. 1995, 409-20.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 36A, SEPTEMBER 2005—2305

You might also like