You are on page 1of 17

12.

10 Austenite Formation and Microstructural Control in Low-Alloy Steels


KD Clarke, Los Alamos National Laboratory, Los Alamos, NM, USA
Ó 2014 Elsevier Ltd. All rights reserved.

12.10.1 Introduction 345


12.10.2 Austenite Formation 345
12.10.3 Typical Heat Treatments and Processing of Austenite 348
12.10.4 Effects of Microstructure, Composition, and Heating Rate on Austenite Formation 352
12.10.4.1 Carbon Steel 352
12.10.4.2 Alloy Steel – Chromium 352
12.10.4.3 Additional Considerations 358
12.10.5 Austenite Structural Development 358
12.10.6 Summary 359
References 359

12.10.1 Introduction

In order to understand steel processing, it is essential to understand the formation and development of austenite (g-iron), the high-
temperature solid face-centered cubic phase. Steel is often heated from room temperature to form austenite multiple times during
industrial processing, as high-temperature austenite requires less energy to deform and shape, and it can be used to effectively
control the final microstructure. In addition, recent developments in advanced high-strength low-carbon sheet steels (AHSS) utilize
novel thermal and deformation processing paths (1), resulting in room temperature-stabilized austenite that can be tailored to
achieve specific mechanical properties tailored to a given application. With a little understanding of the physical metallurgy of steel,
these variations in processing also allow the microstructure, and therefore properties, of a given piece of steel to be varied signif-
icantly solely with controlled variation in mechanical and thermal processing. The results of heat treatment and deformation are
critically dependent on the condition of the prior austenite, even though many low-alloy steels do not contain much, if any,
austenite in service. The wide variety of ways that the condition of the austenite can be controlled and manipulated during pro-
cessing is a significant reason for the broad ranges of steel properties available with very small changes in composition. This
flexibility makes steel the most important structural material in the modern world.
Many excellent resources are available that discuss the details of austenitizing thermal treatments (2–18), including introductory
and specialized textbooks, reference texts, and conference proceedings. This chapter will focus on common austenitizing heat
treatments and the details of austenite formation from various room temperature microstructures, including references to recent
studies. It is important to understand how austenite forms in order to understand the motives behind common austenitizing heat
treatments; these differences in austenite condition that are available to the metallurgist then allow for control of the final
microstructure.

12.10.2 Austenite Formation

For heat treatment of steels, the first resource to become familiar with is the iron–cementite equilibrium phase diagram, which
shows the equilibrium phases in iron–carbon alloys for a given temperature and composition. The iron–carbon equilibrium phase
diagram (10) presented in Figure 1 shows carbon levels up to 7 wt.%, but steels are iron–carbon alloys only up to approximately
2 wt.%, which is the limit of carbon solubility in austenite. The dashed lines show iron–graphite equilibrium phase stability, which
requires extremely long times to attain at low temperatures and carbon levels, and is primarily of interest for cast irons, which have
greater than 2 wt.% carbon. The solid lines indicate metastable iron–cementite (Fe3C, or carbide) equilibrium, which is referred to
for all practical steel-processing considerations.
Steels are generally classified by carbon content, with hypoeutectoid (below 0.77 wt.% carbon), eutectoid (at 0.77 wt.% carbon),
or hypereutectoid (above 0.77 wt.% carbon) steels, each of which has a solid solution of carbon in austenite at high temperature.
Below the A1 temperature of 727  C (referred to as the eutectoid or lower critical temperature), the equilibrium mixture is body-
centered cubic ferrite (a-iron) and cementite. Note that various values are reported for the eutectoid composition and temperature,
varying from 0.76 to 0.83 wt.% carbon and from 722 to 732  C, but consensus-accepted values are 0.76–0.77 wt.% carbon and
727  C, respectively (10,19). However, a binary Fe–C alloy without any impurities is rarely considered, and alloying changes vary
the eutectoid composition and temperature significantly, so exact values are somewhat impractical. Phase stability changes as
a function of composition are discussed in this chapter.
For hypoeutectoid steels, a phase field of ferrite and austenite is stable up to the A3 temperature. This phase field is commonly
referred to as the ‘intercritical’ phase field, because it occurs between the lower and upper intercritical temperatures. For hyper-
eutectoid steels, an austenite–cementite phase field occurs between A1 and Acm, and austenite is stable above Acm (both A3 and Acm

Comprehensive Materials Processing, Volume 12 http://dx.doi.org/10.1016/B978-0-08-096532-1.01211-5 345


346 Austenite Formation and Microstructural Control in Low-Alloy Steels

Figure 1 Carbon-rich region of the iron–carbon (solid lines) and iron–graphite (dashed lines) equilibrium phase diagram. Reproduced from Ericsson,
T. Principles of Heat Treating of Steels, Heat Treating. In ASM Handbook, ASM International, 1991; Vol. 4, pp 3–19.

are referred to as the upper critical temperature). Finally, the phase diagram in Figure 1 is also applicable only to equilibrium
conditions (essentially quasistatic heating or cooling rates), and therefore all equilibrium transformation temperatures are generally
denoted with a nonsubscript ‘e,’ such as Ae1, Ae3, and Aecm. Since the transformation temperatures can vary significantly with
increasing heating or cooling rates, common nomenclature for on-heating and on-cooling transformation temperatures is Ac1 and
Ar1, which are derived from the French words for heating (chauffage) and cooling (refroidissement), respectively, and should be
quoted with heating or cooling rate data for practical utility.
Upon heating an iron–carbon alloy of the eutectoid carbon composition, it will begin to transform directly from the room
temperature starting microstructure (pearlite, bainite, martensite, or some combination) to austenite at the eutectoid temperature,
727  C. Hypoeutectoid steels will begin to form austenite but retain ferrite until reaching the A3 temperature. Steels heated within
the ferrite-plus-austenite phase field are said to be intercritically annealed, and this process can be used to create room temperature
microstructures with various volume fractions of ferrite and martensite (transformed from austenite). At present, intercritical
annealing typically concerns sheet steels that can be processed with relatively consistent through-thickness heating and cooling
rates, which result in homogeneous microstructures and are difficult to apply successfully to thick sections where significant
microstructural gradients occur due to heat transfer considerations. Hypereutectoid steels form austenite while retaining cementite
Austenite Formation and Microstructural Control in Low-Alloy Steels 347

up to the Acm temperature and are fully austenitic above Acm. Thus, the carbon composition in a given steel is the first major factor
determining heat treatment temperature, since all steels experience some processing in the austenite, austenite-plus-ferrite, or
austenite-plus-carbide phase field, regardless of carbon content.
Any other chemical additions, whether intentionally added or not, affect equilibrium transformation temperatures and
composition as noted in this chapter. For example, increasing Cr tends to increase the Ae1 temperature (stabilizing ferrite), whereas
increasing Mn decreases the Ae1 temperature (stabilizing austenite). Generally, Mn, Cu, and Ni are austenite stabilizers, whereas Ti,
Mo, W, Al, P, Si, V, and Cr are ferrite stabilizers, and all of these elements other than Al, P, and Si stabilize cementite and form
carbides (10,20). For practical purposes, all alloying elements also tend to lower the eutectoid carbon concentration (10).
Figure 2(a) shows a plot of the effects of individual alloying elements, along with calculated pseudobinary sections of iron–carbon

Figure 2 (a) Effect of individual alloying elements on equilibrium phase stability as a function of average composition in wt.%. Reproduced from
Ericsson, T. Principles of Heat Treating of Steels, Heat Treating. In ASM Handbook, ASM International, 1991; Vol. 4, pp 3–19. (b–d) Calculated changes in
pseudobinary iron–carbon sections of equilibrium phase diagrams with indicated changes in average composition in accordance with 10XX, 51XX, and
52XXX steels, respectively, showing significant changes in phase stability with composition changes in low-alloy steels. Calculations made with
ThermoCalc. Reproduced from Clarke, A.J.; Speer, J.G.; Miller, M.K.; Hackenberg, R.E.; Edmonds, D.V.; Matlock, D.K.; Rizzo, F.C.; Clarke, K.D.; De Moor,
E. Carbon Partitioning to Austenite from Martensite or Bainite during the Quench and Partition (Q&P) Process: A Critical Assessment. Acta Mater. 2008,
56, 16–22.
348 Austenite Formation and Microstructural Control in Low-Alloy Steels

equilibrium phase diagrams for steels with compositions corresponding to typical AISI 10XX, 51XX, and 52XXX steels, shown in
Figure 2(b)–2(d) (21), respectively. Even though the total alloying additions sum to below 3 wt.% for each case, the phase diagrams
change significantly. First, a three-phase field becomes apparent in all cases where ferrite, austenite, and cementite are all stable. The
eutectoid carbon composition decreases with total alloying content, and the addition of chromium in amounts of 0.8 and 1.4 wt.%
appears to be the major factor in increasing the eutectoid transformation temperature by 20  C. Thus, the substitutional alloy
content is another major factor in determining heat treatment temperatures. Generally, appropriate heat treatment temperatures are
available in the Heat Treater’s Guide (13), but new alloys or nonstandard thermal cycles (e.g., induction heating) may require some
adjustment to recommended practice.

12.10.3 Typical Heat Treatments and Processing of Austenite

Common heat treatments, such as full annealing, spheroidizing, normalizing, intercritical annealing, homogenizing, and hard-
ening, are discussed. There are many resources available that cover various processes in greater detail or give specific thermal cycles
for a given steel composition (2–18), but the basics of these thermal treatments are summarized here.
The purpose of full annealing is to erase the previous room temperature microstructure and soften previously strain-hardened
material, generally for ease of subsequent deformation processing or machining. Full annealing is assumed when the term
‘annealing’ is used without precursors. In order to fully anneal hypoeutectoid steel, it must be heat treated to just above the Ae3
temperature. For example, 1015 steel (0.15 wt.% carbon) would be annealed at 900  C, whereas 1045 steel (0.45 wt.% carbon)
would be annealed at 845  C, since the Ae3 transformation temperature decreases with increasing carbon content (13). The material
is held at the annealing temperature, typically for 1 h, and then furnace-cooled slowly. This annealing treatment produces a coarse
ferrite–pearlite microstructure and the lowest strength for a given alloy, which is most useful for subsequent cold deformation
processes.
For hypereutectoid compositions, annealing is performed above the Ae1 temperature, since the preferred microstructure for
subsequent heat treatment or machining is ferrite and spheroidized carbides. For example, 52100 is annealed at 790  C, and the
material must then be cooled very slowly to allow austenite to transform to ferrite, avoid other transformations, and allow carbides
to coarsen (carbon diffusion to carbides). This type of annealing is referred to as a spheroidizing anneal. If it is necessary to further
spheroidize carbides for machinability or deformation processing, subsequent treatments are performed just below the lower
critical temperature, which is used for medium- and high-carbon steels. Spheroidizing treatments can take many hours, but they
may be the only way to effectively machine medium- and high-carbon steels, which can have high strength even with a coarse
pearlite microstructure. For hypereutectoid steels, cycling above and below the lower critical temperature has also been used to
refine the microstructure (22).
Both hypoeutectoid and hypereutectoid steels can be normalized, a process that is used to erase heavily cold-worked micro-
structure or the effects of hot forging. Normalizing creates a fine, equiaxed austenite grain size and a subsequent fine room
temperature microstructural scale that is easier to process or heat treat, and results in improved toughness. Normalizing involves
heating to temperatures above the upper critical temperature, allowing austenite formation to complete, and cooling in air. Slightly
higher austenitizing temperatures and a faster cooling rate are the primary differences between normalizing and full annealing, and
they result in a fine, equiaxed, room temperature grain size. It should be noted that normalizing is often applied to parts forged at
very high temperatures where austenite grain size may be very coarse. Austenite grain growth kinetics increase with temperature, and
rapid grain growth occurs above approximately 1050  C, even in killed steels with fine AlN precipitates (Figure 3(a) and 3(b))
(17,18). These precipitates control austenite grain growth until the dissolution temperature is reached, which coincides with the
grain growth temperature (Figure 3(c)) (3). Since forging can require very high temperatures, a subsequent normalizing treatment is
used to refine the austenite grain size and room temperature ferrite–pearlite microstructure.
As mentioned in this chapter, intercritical annealing treatments can be used to create microstructures with fixed ferrite and
austenite volume fractions, resulting in controlled volume fractions of ferrite and martensite and/or bainite. Intercritically annealed
steels are generally quenched from the two-phase field, and the quench is commonly the final heat treatment before the steel is put
into service. This practice results in products with excellent strength and ductility combinations, and is used to produce dual-phase
(DP) steel, for example. For this reason, intercritical annealing is an extremely important process in the production of AHSS low-
carbon sheets, and knowledge of precise phase stability and control of alloy content is necessary.
Homogenizing heat treatments are used to reduce the elemental segregation that occurs during casting of steels. Steels solidify by
the growth of dendritic crystals, which reject alloy and impurity elements into the liquid according to equilibrium phase solubility
requirements, resulting in the phenomenon of interdendritic segregation (3), which causes both macro- and microsegregation.
Macrosegregation refers to gradual changes in average composition over large distances (centimeters to meters) as solidification
progresses, and it can require extremely long times (days) at high temperatures to erase, which is not generally a practical remedy
during industrial processing. Microsegregation refers to localized concentration gradients that are a product of solidification and
that span distances of millimeters or less. Dendritic segregation is a common example, which results in banded microstructures in
rolled material. An example of microstructurally banded steel is shown in Figure 4 (21); this results from chromium segregation in
hot-rolled 52100 steel. The banding scale is on the order of 50 microns and varies from 1.2 to 2.0 wt.% chromium. It is evident that
this type of composition variation can affect local phase stability, as shown by the calculated equilibrium phase fractions in
Figure 4. Removing this type of segregation still requires long heat treatment times at high temperatures. Based on Figure 5 (21,23),
Austenite Formation and Microstructural Control in Low-Alloy Steels 349

Figure 3 Austenite grain growth in (a) Ck 15 and (b) Ck 45 steel as a function of heating rate and maximum temperature. Aluminum and nitrogen
contents in the Ck 15 steel are 0.006 and 0.0037, respectively (semikilled). Aluminum and nitrogen contents in the Ck 45 steel are 0.014 and
0.0115, respectively (killed). Reproduced from Orlich, J.; Rose, A.; Wiest, P. Atlas zur Wärmebehandlung der Stähle, Band 3; Verlag Stahleisen M.B.H.:
Düsseldorf, Germany, 1973 and Orlich, J.; Pietrzeniuk, H.-J. Atlas zur Wärmebehandlung der Stähle, Band 4; Verlag Stahleisen M.B.H.: Düsseldorf,
Germany, 1976. (c) Generalized austenite grain size as a function of temperature for slow heating rates, showing grain-coarsening temperature for
fine-grained, killed steels (approx. 1050  C), indicating the dissolution temperature of AlN precipitates, and gradual austenite grain growth in coarse-
grained steels. Reproduced from Krauss, G. Steels Processing, Structure, and Performance; ASM International: Materials Park, OH, 2005.
350 Austenite Formation and Microstructural Control in Low-Alloy Steels

Figure 4 (a) Optical micrograph of spheroidized 52100 steel showing microstructural blanding parallel to the rolling direction (arrow), nital etch. (b)
Microprobe results of chromium concentration variation in the steel depicted in (a) over 300 mm. Equilibrium phase content during (c) ferrite-to-
austenite and (d) carbide dissolution based on calculated phase stability with variations in chromium content equivalent to those measured in the
microprobe analysis in (b). Calculations made in ThermoCalc. Reproduced from Clarke, K.D. The Effect of Heating Rate and Microstructural Scale on
Austenite Formation, Austenite Homogenization, and As-quenched Microstructure in Three Induction Hardenable Steels. Ph.D. Thesis, Colorado
School of Mines: Golden, CO, 2008.

chromium can diffuse 10 microns in 1000 s at 1200  C, so eliminating this segregation would take a minimum of 5000 s
(approximately 1.5 h), which would not be desirable, as it would also greatly increase the austenite grain size. However, it has been
shown that real homogenization times take longer than diffusion estimates would predict, primarily due to the decreasing driving
force as segregation is reduced. In plain-carbon steels, manganese banding is often found in hot-rolled steels, and a study (24)
examining the homogenization times required to remove 95% of the manganese segregation in hot-rolled 4145 steel found that
100 h at 1200  C are needed. Chemically banded microstructures are of particular concern when considering surface-hardened
components that are heat treated using high-heating-rate methods (induction or flame hardening), or heat treatments that use
intercritical annealing and local phase stability control to create complex final microstructures, such as those for AHSS (1,25–27).
Hardening heat treatments in steels are used to create martensitic or perhaps bainitic microstructures, and they require fast
cooling rates from austenite to avoid transformation into pearlite. Figure 6 (10) shows a continuous cooling diagram for steel
containing 0.45 wt.% carbon. In order to avoid forming pearlite, a very fast quench rate is necessary from the austenitizing
temperature, with the temperature of the material having to cool to below 500  C within a few seconds. For this alloy, the hard-
enability is low enough (i.e., required quench rates are very fast) that it is difficult to achieve the desired quench rate to get a fully
martensitic microstructure. Increasing alloy content or manipulating grain size can be used to increase hardenability (3).
Austenite Formation and Microstructural Control in Low-Alloy Steels 351

Figure 5 Diffusion distance in austenite at isothermal holds for indicated times, estimated from L ¼ 2ODt, where L represents the diffusion distance in
centimeters, D is the diffusion coefficient of the diffusing species in austenite (cm2 s1), and t is time in seconds. Reproduced from Clarke, K.D.
The Effect of Heating Rate and Microstructural Scale on Austenite Formation, Austenite Homogenization, and As-quenched Microstructure in Three
Induction Hardenable Steels. Ph.D. Thesis, Colorado School of Mines: Golden, CO, 2008. Diffusion coefficients from: Shewmon, P. Diffusion in Solids,
2nd ed.; TMS: Warrendale, PA, 1989.

Figure 6 Continuous cooling transformation diagram for 1045 steel, indicating the cooling rates required to achieve specified final microstructures.
Reproduced from Ericsson, T. Principles of Heat Treating of Steels, Heat Treating. In ASM Handbook; ASM International, 1991; Vol. 4, pp 3–19.
352 Austenite Formation and Microstructural Control in Low-Alloy Steels

12.10.4 Effects of Microstructure, Composition, and Heating Rate on Austenite Formation

In addition to the equilibrium temperature at which austenite becomes stable for a given steel composition, the rate at which
austenite forms (kinetics) also varies with initial microstructure, composition, austenitizing temperature, and heating rate. The
kinetics of a given reaction means that the simple equilibrium iron–cementite phase diagram must be supplemented with diagrams
incorporating composition, microstructure, and the rate at which reactions take place, such as the continuous-heating time–
temperature–austenitization (TTA) diagram. An excellent resource for a wide selection of TTA diagrams is the Atlas zur Wärmebe-
handlung der Stähle (15–18), an encyclopedic four-volume resource produced between 1961 and 1976 that includes continuous-
heating and isothermal-hold TTA diagrams for most steel compositions in common use at the time, in addition to excellent
data on recommended heat treatment, effects of chemistry variations, hardenability, isothermal austenite decomposition,
continuous cooling transformation, carburizing, austenite grain size, as-quenched hardness, martensite start temperatures, and
micrographs of resulting microstructures.
Here, we focus on the formation of austenite by examining continuous-heating TTA diagrams for plain carbon and chromium
alloy steels with various starting microstructures, as shown in Figures 7–11. This series demonstrates the effects of the heating rate,
starting microstructure, and composition. The examples focus on chromium substitutional alloying, but the concepts also apply to
other solid-solution alloying elements.

12.10.4.1 Carbon Steel


Figure 7 shows continuous-heating TTA diagrams for Ck 15 and Ck 45 steels (equivalent to AISI 1015 and 1045), each starting from
ferrite–pearlite initial microstructures (17,18).
The TTA diagram shows that the initial transformation from ferrite–pearlite to austenite (Ac1) occurs at much higher temper-
atures with increasing heating rate, increasing from approximately 730  C to nearly 790  C. At high enough temperatures, austenite
can be stable without carbon (911  C in Figure 1), but at these lower temperatures austenite requires dissolved carbon to be stable.
On heating, therefore, austenite nucleates at the carbides, which supply carbon to the growing austenite grains. Carbon must also
diffuse from the carbide through the austenite to the growth front in order to allow further austenite growth, resulting in diffusion-
controlled growth kinetics. Thus, the carbide dissolution and carbon diffusion kinetics are critical to austenite formation kinetics. In
the case of a ferrite–pearlite microstructure, nucleation occurs at pearlite grain boundaries (3). For the continuous-heating TTA
shown in Figure 7, it is also important to recognize that the ferrite-plus-pearlite-plus-austenite temperature range for the 1045 steel
also increases with the increasing heating rate, which, as is shown in this chapter, is an effect of a coarse initial microstructure.
Finally, in order to achieve homogeneous austenite, with respect to carbon, even higher temperatures are required, which can be
very important to the subsequent transformations that occur when cooling this austenite to room temperature. Homogenization of
substitutional alloying element segregation, as discussed in this chapter, requires very long times.
The effects of carbon content are also evident in Figure 7, with the Ac3 temperature for the lower carbon steel significantly
higher than for the 1045 steel, and thus requiring significantly higher austenitization temperature (approximately 860  C vs
770  C) for slow heating rates. The recommended standard austenitizing temperatures for annealing or normalizing these grades
are similar – respectively 885 and 915  C for 1015, and 845 and 900  C for 1045 (13) – but do show that the 1015
composition requires higher temperatures for austenitizing. However, when heating rates are increased much higher than
100  C s1, the transformation temperatures are nearly equal, due to the kinetics of carbide dissolution and carbon diffusion
required to form austenite: for the 1015 microstructure, the carbon diffusion distance is greater, since there is more ferrite,
whereas for the 1045 microstructure, the larger volume fraction of carbides that must dissolve is likely the primary process
controlling the rate of full austenitization. It is also important to note that the temperature at which homogeneous (i.e., no
carbon gradients) austenite occurs is similar at high heating rates. The rate of carbon diffusion, as an interstitial element, is fast
relative to substitutional elements, but heating rates above 1000  C s1 are fast enough so that it must be considered a rate-
limiting process; see Figure 5 (21,23). The starting microstructure in Figure 7 also shows significantly spheroidized pearlite
for the 1045 starting microstructure, which suggests that the dissolution kinetics may be sluggish relative to a starting micro-
structure of pearlite with fine lamellae. Several studies have been performed on the effect of microstructural scale on 1045
austenitization kinetics (21,28–30). One study (21,28) compared a hot-rolled microstructure with 30 vol.% ferrite and a larger
pearlite colony size with a normalized microstructure with 40 vol.% ferrite and a smaller pearlite colony size. Results are shown
in Figure 8. Although the apparent effect on the transformation temperatures is small, the change in as-quenched hardness is
dramatic: the normalized starting microstructure responds to austenitization heat treatment much more quickly than the coarser
hot-rolled microstructure. These results show that, for industrial nonequilibrium heating or cooling rate processing, the use of
a normalizing treatment to refine grain size is beneficial to achieve a faster austenitization response, which requires shorter heat
treatments and lower maximum temperatures.

12.10.4.2 Alloy Steel – Chromium


When additional alloying and variations in microstructure are considered, the TTA diagram can vary significantly. Here, we consider
chromium alloying effects on transformation behavior for a 46Cr2 composition with a ferrite–pearlite starting microstructure and
a 41Cr4 composition with a tempered martensite starting microstructure, as shown in Figure 9 (17,18). Both compositions fall
Austenite Formation and Microstructural Control in Low-Alloy Steels 353

Figure 7 Continuous-heating time–temperature–austenitization (TTA) diagrams and optical micrographs of starting microstructures for Ck 15 (left)
and Ck 45 (right) steels with ferrite–pearlite starting microstructures. Note that the pearlite for the Ck 45 steel is partially spheroidized, resulting in large
carbides. Reproduced from Orlich, J.; Rose, A.; Wiest, P. Atlas zur Wärmebehandlung der Stähle, Band 3; Verlag Stahleisen M.B.H.: Düsseldorf,
Germany, 1973 and Orlich, J.; Pietrzeniuk, H.-J. Atlas zur Wärmebehandlung der Stähle, Band 4; Verlag Stahleisen M.B.H.: Düsseldorf, Germany, 1976.
354 Austenite Formation and Microstructural Control in Low-Alloy Steels

Figure 8 Measured changes in on-heating transformation temperatures for hot-rolled and normalized 1045 steel (top), along with optical micrographs
of starting microstructures and final hardness for each microstructure (middle row: hot-rolled; bottom row: normalized). Reproduced from Clarke,
K.D. The Effect of Heating Rate and Microstructural Scale on Austenite Formation, Austenite Homogenization, and As-quenched Microstructure in Three
Induction Hardenable Steels. Ph.D. Thesis, Colorado School of Mines: Golden, CO, 2008 and Lee, S.-J.; Clarke, K.D.; Van Tyne, C.J. An On-Heating
Dilation Conversional Model for Austenite Formation in Hypoeutectoid Steels. Metall. Mater. Trans. A September 2010, 41 (9), 2224–2235.
Austenite Formation and Microstructural Control in Low-Alloy Steels 355

Figure 9 Continuous-heating time diagrams for (a) 46Cr2 and (b) 41Cr4 steels with optical micrographs of starting microstructures of (a) ferrite–
pearlite, and (b) martensite. Compositions are similar to that of AISI 5150. Reproduced from Orlich, J.; Rose, A.; Wiest, P. Atlas zur Wärmebehandlung der
Stähle, Band 3; Verlag Stahleisen M.B.H.: Düsseldorf, Germany, 1973 and Orlich, J.; Pietrzeniuk, H.-J. Atlas zur Wärmebehandlung der Stähle, Band 4;
Verlag Stahleisen M.B.H.: Düsseldorf, Germany, 1976.

within values for AISI 5150H alloy, with the exception of a low-carbon content (0.4 wt.%) and minor Mo addition (0.04 wt.%) in
41Cr4. Standard annealing and normalizing temperatures for 5150 are 830 and 870  C, which are just above the Ac3 temperatures
shown in Figure 9.
In the example given here of carbon steels (Figures 7 and 8), the effect of microstructure scale was difficult to determine, since the
level of carbon was significantly different (Figure 7) or the effect was subtle (Figure 8). For the example shown in Figure 9, however,
the effects of coarser microstructures are evident, with an increasing scale of microstructure increasing the effect of the heating rate,
because austenite formation is a nucleation and growth process. For chromium-alloyed steels in particular, the dissolution of the
resultant chromium-enriched carbides is slower than for carbides in plain-carbon steels. At slow heating rates, the effect is relatively
small, but as heating rates increase, the change in transformation temperatures is much greater for the ferrite–pearlite starting
structure, even though the material chemistry is very similar. In fact, carbides are still undissolved in the ferrite–pearlite initial
microstructure above 1000  C when considering the highest heating rates. In a study of induction-heated specimens of 5150 steel,
356 Austenite Formation and Microstructural Control in Low-Alloy Steels

Figure 10 (a) Scanning electron micrograph showing evidence of retained carbides for a 5150 hot-rolled starting microstructure continuously
heated to 900  C at 300  C s1 and immediately quenched. The measured hardness and dilatometric analysis suggest full austenitization. Reproduced
from Clarke, K.D. The Effect of Heating Rate and Microstructural Scale on Austenite Formation, Austenite Homogenization, and As-Quenched Micro-
structure in Three Induction Hardenable Steels. Ph.D. Thesis, Colorado School of Mines: Golden, CO, 2008; Clarke, K.D.; Van Tyne, C.J. The Effect of
Heating Rate and Prior Microstructure on Austenitization Kinetics of 5150 Hot-Rolled and Quenched and Tempered Steel. In Materials Science &
Technology 2007; TMS: Detroit, MI, 2007 and K.D. Clarke, Van Tyne, C.J.; Hackenberg, R.E.; Vigil, C.J. Induction Hardening 5150 Steel – Effects of Initial
Microstructure and Heating Rate, winner of the 2010 HTS/Bodycote Best Paper in Heat Treating Award, March 2, 2010, announced in Adv. Mater.
Processes, August, 2010. Published in the J. Mater. Eng. Perform. 2011, 20 (2),161–168; also published electronically in ASM Heat Treating Society
Industrial Heat Treating and HTPro eNews April 6, 2010. (b, c) Dictra diffusion calculations showing calculated diffusion distances of carbon, chromium,
and manganese from (b) austenite into ferrite (simulating transformation of proeutectoid ferrite after pearlite regions have transformed to austenite),
and (c) carbide into austenite (simulating carbide dissolution in an austenite field). (b) Suggests very little substitutional diffusion into prior ferrite
regions. (c) Shows carbide dissolution is dependent on the diffusion of substitutional alloy away from the carbide. Reproduced from Clarke, K.D.
The Effect of Heating Rate and Microstructural Scale on Austenite Formation, Austenite Homogenization, and As-Quenched Microstructure in Three
Induction Hardenable Steels. Ph.D. Thesis, Colorado School of Mines: Golden, CO, 2008.

it was found that evidence of prior pearlite lamellae was obvious in the microstructure at temperatures well above Ac3 (21,31,32)
(see Figure 10(a)), although maximum as-quenched hardness was achieved. These carbide remnants have been referred to as ‘ghost
pearlite’ (33) and were not found in 1045 steels with similar starting microstructures and high-rate heat treatments (21). This, along
with hardness values that indicate most carbon is in solution, suggests that these remnants of pearlite may simply be locally high
chromium concentrations, which result from partitioning of chromium to pearlitic carbides and slow diffusion of chromium
relative to carbon, as shown in Figure 5. Metallographic evidence of chromium partitioning is evident, since there is a significant
etching response. However, other substitutional elements, such as manganese, may also partition to carbides but would not be
Austenite Formation and Microstructural Control in Low-Alloy Steels 357

Figure 11 (a) Scanning electron micrographs of prior-spheroidized 52100 steel heated at 300  C s1 to 1200  C, and the resulting evidence of alloy
partitioning to locations of prior carbides. (b) Dictra diffusion calculations showing that significant chromium and manganese segregation remains
after carbide dissolution (after 1.2 s). Reproduced from Clarke, A.J.; Speer, J.G.; Miller, M.K.; Hackenberg, R.E.; Edmonds, D.V.; Matlock, D.K.;
Rizzo, F.C.; Clarke, K.D.; De Moor, E. Carbon Partitioning to Austenite from Martensite or Bainite during the Quench and Partition (Q&P) Process:
A Critical Assessment. Acta Mater. 2008, 56, 16–22.

evident metallographically since manganese does not reduce etch attack significantly like chromium. Therefore, the Dictra simu-
lations presented in Figure 10 (21) also suggest that chromium and manganese diffusion from substitutional alloy-partitioned
carbides slows carbide dissolution during high-rate, short-time heat treatments. In addition, substitutional diffusion to prior
ferrite regions results in significant concentration gradients. Although these locally high concentrations occur over very short
distances, there may be subsequent effects on the local phase stability, microstructure formation, mechanical and fatigue properties,
or corrosion resistance of materials with this remnant chromium or manganese segregation.
A further example of austenite formation kinetics with chromium alloying is 52100 steel, a hypereutectoid carbon composition
with nominally 1.5 wt.% chromium, which is generally heat treated from a heavily spheroidized carbide condition, as shown in
Figure 11, that requires significant time for dissolution (21,34–36). After heating at 300  C s1 to 1200  C, remnant carbides
remain, but regions where carbides fully dissolved are also evident, likely because of remnant alloy originally partitioned to the
carbide. The accompanying Dictra calculations show that maximum temperatures above 1400  C for high heating rates would still
not eliminate the substitutional alloy segregation to the locations of prior carbides. Again, this suggests that high-rate heating results
in chemically inhomogeneous microstructures, and there is every reason to suspect that the limited diffusivity of substitutional
species results in local inhomogeneity even during slower heat treatment.
358 Austenite Formation and Microstructural Control in Low-Alloy Steels

12.10.4.3 Additional Considerations


A significant number of articles have been published on the formation of austenite (21,28–99) over many years. However, there
appears to be renewed interest in austenite formation over the past decade because of the new processing routes that are either in use or
being considered for first- and second-generation AHSS (low-carbon sheet) applications (1), which include DP, complex-phase,
transformation-induced plasticity (TRIP), lightweight steels with induced plasticity (L-IP), shear band strengthened, and twinning-
induced plasticity (TWIP). In addition, research is ongoing to determine the viability of enhanced DP steels, modified TRIP, ultrafine-
grained bainite, quenching and partitioning (Q&P), flash processing, lower manganese TWIP and TRIP, and high manganese TRIP as
the third generation of AHSS, where it is expected that final microstructures consisting of ferrite, austenite, and martensite will be
required to meet strength and ductility requirements (1). Although the specifics of each process will not be discussed in detail here, it is
obvious that composition and microstructure are the key variables that can be selectively controlled to produce product properties to
meet industry requirements. The addition of higher quantities of substitutional alloying elements such as manganese, chromium,
silicon, aluminum, cobalt, nickel, and others affects the phase stability of austenite, and this must be considered during processing.
Controlling grain size and incorporating steels that deform via twinning rather than slip also pose challenges for processing austenite
and affect the formation and development of austenite. Finally, these new processing paths may also include deformation of the
austenite to mechanically decrease grain size, as microalloying approaches may no longer be compatible with alloying strategies.
Intercritical heat treatments used to tailor specific volume fractions of ferrite and martensite and/or bainite should also be
examined. During intercritical annealing, austenite forms with carbon content that is specified by a horizontal tie line between the
A3 and the ferrite phase field (essentially zero carbon). The lever rule can be used to calculate phase fraction (2), but the intercritical
temperature used determines the austenite carbon content. Obviously, small variations in temperature and phase stability for these
heat treatments will result in large changes in phase fraction and austenite carbon content, and thus precise understanding of the
equilibrium phase stability for the particular composition is vital, along with accurate process control. Using intercritical annealing
can result in ferrite-plus-martensite microstructures, or can be designed to create austenite with high carbon or other alloy levels,
which can stabilize austenite to room temperature, assuming no competing phases (carbides) form during cooling. Alloying with
aluminum or silicon can impede carbide growth. Additionally, an important consideration when alloying with significant amounts
of aluminum is that full austenitization may not be possible.
In addition to composition, heating rate, and microstructural-scale effects, residual driving force within the microstructure can
affect austenitization. Cold work and untempered martensite are two examples of potential energy stored within the microstructure.
Many modern steel-processing paths utilize intercritical heat treatment in order to tailor microstructure with optimized volumes of
ferrite and martensite (or bainite) to design steels with high strength and good ductility. Recent advances in high-strength steels may
also require postquenching heat treatments in microstructures that retain austenite to room temperature via carbon stabilization
and the prevention of carbide formation. The local stability is therefore critically important to the final properties.
Cold-worked materials tend to recrystallize before austenite formation in standard furnace heat treatments. However, it has been
shown that heating rates above some critical value (97–99) result in a reduction in the required hold times to achieve a given
amount of austenite formation, as recrystallization during austenitization increases the rate of austenite formation. It would follow
that untempered martensite microstructures may alter austenite formation kinetics if a critical heating rate is used to limit tempering
before entering the austenite phase field, if tempering and austenitization can occur simultaneously.

12.10.5 Austenite Structural Development

Once austenite formation is complete, the condition of austenite as a function of heat treatment is of critical importance to the final
properties of the steel. In particular, austenite grain size has an important effect on subsequent processing. The temperatures and
times at which steels are heat treated in the austenite phase field can have significant effects on austenite grain size. Several ways to
control austenite grain size include using aluminum additions, microalloying, and thermomechanical processing (3,4,6,8,9).
Aluminum and nitrogen additions have been shown to minimize austenite grain growth up to 1000  C in standard furnace heat
treatments (3). As shown in Figure 3, differences in austenite grain growth occur as a function of aluminum content for contin-
uously heated steels. The Ck 15 steel has 0.06 wt.% aluminum (semikilled), whereas the Ck 45 steel has 0.14 wt.% aluminum
(killed), and their austenite grain growth rates are substantially different, especially when considering slow heating rates. Very
specific quantities of aluminum and nitrogen are required to form aluminum nitride (AlN) particles that pin austenite grain
boundaries up to 1000  C. If the aluminum content is below 0.08 wt.%, there are not enough particles to pin grain boundaries,
whereas if the aluminum content is above 0.15 wt.%, the particles tend to coarsen and be ineffective in pinning grain boundaries.
Note that above 1000  C for slow heating rates, the aluminum nitride precipitates dissolve, and very similar grain sizes are found for
both steels. At higher heating rates, it is interesting to note that the dissolution kinetics of aluminum nitrides results in small grain
sizes in the fully aluminum-killed steel up to very high temperatures (over 1100  C at 100  C s1). It is therefore possible to use
rapid-heating techniques up to higher maximum temperatures than would ordinarily be considered, as long as hold times at
a temperature are minimized or eliminated.
Other alloying elements, such as titanium, vanadium, and niobium, are often used in low-alloy steels as microalloying additions
to control austenite grain size. These microalloying additions perform a similar function to aluminum in killed steels, but are used
in conjunction with nitrogen or carbon to control the stability and solubility of the desired precipitate (3).
Austenite Formation and Microstructural Control in Low-Alloy Steels 359

As mentioned in this chapter, high-temperature austenite has greatly reduced resistance to deformation, and it is therefore
favorable to deform steels in the austenite phase field whenever possible. Thermomechanical processing of steels in the austenite
phase field, in conjunction with controlled recrystallization, can also be used to minimize austenite grain size and therefore also the
final grain size in the transformed microstructure. In microalloyed steels, austenite recrystallization may be suppressed by
precipitation of fine microalloy carbonitride particles, resulting in grain structures elongated in the rolling direction. This process is
known as controlled rolling. More detailed reviews of thermomechanical processing in steels are available in the literature, for
example Refs. (3,4,8,9).

12.10.6 Summary

Austenite plays a key role in the processing of steels, and the condition of austenite that is formed during processing can be
controlled to optimize the subsequent in-service microstructure. Many austenitization heat treatments have been optimized over
decades for specific alloy compositions and processing capabilities. However, there is no ‘ideal’ austenitic microstructure for every
application. It is generally advantageous to minimize austenite grain growth during processing, while ensuring the alloy content in
the microstructure is as homogeneous as possible. Grain growth can be controlled by controlling processing temperatures,
microalloying, or thermomechanical processing. Austenite composition and homogeneity should be carefully considered for high-
heating-rate processes (e.g., surface induction hardening), and they may become more significant with modern developments in
low-carbon AHSS sheets, which require precise processing paths and a thorough understanding of local phase stability.
The breadth of available austenite microstructures, however, affords the steel metallurgist a wide variety of processing oppor-
tunities that are not available in other alloy systems. This allows steel to persist as the most important structural material in the
modern world, and it has resulted in major advancements in steel applications continuously for over a century. Recent improve-
ments in process capabilities and theoretical understanding in steels suggest this trend will continue into the foreseeable future.

See also: Mathematical Modeling of Weld Phenomena. Part 1: Finite-Element Modeling; Bonding Technologies in Manufacturing
Engineering; Review of Microstructures, Mechanical Properties, and Residual Stresses of Ferritic and Martensitic Stainless-Steel
Welded Joints; Advances in Fused Deposition Modeling; Laser-Assisted Additive Manufacturing for Metallic Biomedical Scaffolds.

References

1. DeMoor, E.; Gibbs, P. J.; Speer, J. G.; Matlock, D. K. Strategies for Third-Generation Advanced High-Strength Steel Development. AIST Trans. 2010, 7, 133–144.
2. Callister, W. D.; Rethwisch, D. G. Materials Science and Engineering, an Introduction, 8th ed.; Wiley and Sons: New York, NY, 2010.
3. Krauss, G. Steels Processing, Structure, and Performance; ASM International: Materials Park, OH, 2005.
4. Brooks, C. R. Principles of the Heat Treatment of Plain Carbon and Low Alloy Steels; ASM International: Materials Park, OH, 1996.
5. Brooks, C. R. Principles of the Surface Treatment of Steels; Technomic: Lancaster, PA, 1992.
6. Brooks, C. R. Heat Treatment of Ferrous Alloys; McGraw-Hill: New York, NY, 1979.
7. Abbaschian, R.; Abbaschian, L.; Reed-Hill, R. E. Physical Metallurgy Principles, 4th ed.; Cengage Learning: Stamford, CT, 2008.
8. Leslie, W. C. The Physical Metallurgy of Steels; Hemisphere: Marietta, OH, 1981.
9. De Cooman, B. C.; Speer, J. G. Fundamentals of Steel Product Metallurgy; AIST, 2011.
10. Ericsson, T. Principles of Heat Treating of Steels, Heat Treating. In ASM Handbook; ASM International, 1991; Vol. 4, pp 3–19.
11. Rajan, T. V.; Sharma, C. P.; Sharma, A. Heat Treatment: Principles and Techniques; Prentice-Hall: New Delhi, India, 2006.
12. Grossman, M. A.; Bain, E. C. Principles of Heat Treatment, 5th ed.; American Society for Metals: Metals Park, OH, 1964.
13. Unterweiser, P. M. Heat Treater’s Guide Standard Practices and Procedures for Steel; American Society for Metals: Metals Park, OH, 1982.
14. Damm, E. B.; Merwin, M. J. Austenite Formation and Decomposition; TMS: Warrendale, PA, 2003.
15. Wever, F.; Rose, A.; Peter, W.; Strassburg, W.; Rademacher, L. Atlas zur Wärmebehandlung der Stähle, Band 1; Verlag Stahleisen M.B.H: Düsseldorf, Germany, 1961.
16. Rose, A.; Hougardy, H. Atlas zur Wärmebehandlung der Stähle, Band 2; Verlag Stahleisen M.B.H: Düsseldorf, Germany, 1972.
17. Orlich, J.; Rose, A.; Wiest, P. Atlas zur Wärmebehandlung der Stähle, Band 3; Verlag Stahleisen M.B.H: Düsseldorf, Germany, 1973.
18. Orlich, J.; Pietrzeniuk, H.-J. Atlas zur Wärmebehandlung der Stähle, Band 4; Verlag Stahleisen M.B.H: Düsseldorf, Germany, 1976.
19. C (Carbon) Binary Alloy Phase Diagrams. In Alloy Phase Diagrams; ASM Handbook; ASM International, 1992,; Vol. 3, pp 2.109–2.116.
20. DeMeyer, M.; Vanderschueren, D.; DeCoomen, B. C. The Influence of the Substitution of Si by Al on the Properties of Cold Rolled C–Mn–Si TRIP Steels. ISIJ Int. 1999, 39 (8),
813–822.
21. Clarke, K. D. The Effect of Heating Rate and Microstructural Scale on Austenite Formation, Austenite Homogenization, and As-quenched Microstructure in Three Induction
Hardenable Steels. Ph.D. Thesis, Colorado School of Mines: Golden, CO, 2008.
22. Kayali, E. S.; Sunada, H.; Oyama, T.; Wadsworth, J.; Sherby, O. D. The Development of Fine Structure Superplasticity in Cast Ultrahigh Carbon Steels through Thermal Cycling.
J. Mater. Sci. 1979, 14, 2688–2692.
23. Shewmon, P. Diffusion in Solids, 2nd ed.; TMS: Warrendale, PA, 1989.
24. Anderson, P.I. Induction Hardening Response of Ferrite and Pearlite Banded Steel. M.S. Thesis, Colorado School of Mines: Golden, CO, 2005.
25. Clarke, A. J.; Speer, J. G.; Miller, M. K.; Hackenberg, R. E.; Edmonds, D. V.; Matlock, D. K.; Rizzo, F. C.; Clarke, K. D.; De Moor, E. Carbon Partitioning to Austenite from
Martensite or Bainite during the Quench and Partition (Q&P) Process: A Critical Assessment. Acta Mater. 2008, 56, 16–22.
26. Clarke, A. J.; Speer, J. G.; Matlock, D. K.; Rizzo, F. C.; Edmonds, D. V.; Santofimia, M. J. Influence of Carbon Partitioning Kinetics on Final Austenite Fraction during Quenching
and Partitioning. Scr. Mater. 2009, 61, 149–152.
27. Gibbs, P. J.; De Moor, E.; Merwin, M. J.; Clausen, B.; Speer, J. G.; Matlock, D. K. Austenite Stability Effects on Tensile Behavior of Manganese-Enriched Austenite
Transformation-Induced Plasticity Steel. Metall. Mater. Trans. A 2011, 42, 3691–3702.
28. Lee, S.-J.; Clarke, K. D.; Van Tyne, C. J. An On-Heating Dilation Conversional Model for Austenite Formation in Hypoeutectoid Steels. Metall. Mater. Trans. A September 2010,
41 (9), 2224–2235.
360 Austenite Formation and Microstructural Control in Low-Alloy Steels

29. Misaka, Y.; Kiyosawa, Y.; Kawasaki, K.; Yamazaki, T.; Silverthorne, W. O. Gear Contour Hardening by Micropulse Induction Heating System. document 970971; Society of
Automotive Engineers, 1997; pp 121–130.
30. López, V.; Fernández, J. M.; Belló, J. M.; Ruiz, J.; Zubiri, F. Influence of Previous Structure on Laser Surface Hardening of AISI 1045 Steel. ISIJ Int. 1995, 35 (11), 1394–1399.
31. Clarke, K. D.; Van Tyne, C. J. The Effect of Heating Rate and Prior Microstructure on Austenitization Kinetics of 5150 Hot-rolled and Quenched and Tempered Steel. In Materials
Science & Technology 2007; TMS: Detroit, MI, 2007.
32. Clarke, K. D.; Van Tyne, C. J.; Hackenberg, R. E.; Vigil, C. J. Induction Hardening 5150 Steel – Effects of Initial Microstructure and Heating Rate. J. Mater. Eng. and Perform.
2011, 20 (2), 161–168.
33. Medlin, D. J.; Krauss, G.; Thompson, S. W. Induction Hardening Response of 1550 and 5150 Steels with Similar Prior Microstructures. In 1st International Conference on
Induction Hardening of Gears and Critical Components. Indianapolis, IN; 1995.
34. Clarke, K. D.; Van Tyne, C. J. Effect of Prior Microstructure Scale and Heating Rate on Carbide Dissolution and Austenite Formation in Spheroidized 52100 Steel. In 27th
Forging Industry Technical Conference and Energy Summit; Forging Industry Association: Ft. Worth, TX, 2007.
35. Lee, S.-J.; Clarke, K. D. A Conversional Model for Austenite Formation in Hypereutectoid Steels. Metall. Mater. Trans. A Dec. 2010, 41 (12), 3027–3031.
36. Lee, S.-J.; Clarke, K. D. A Quantitative Investigation of Cementite Dissolution Kinetics for Continuous Heating of Hypereutectoid Steel. Metall. Mater. Trans. A, submitted
for publication.
37. Bénéteau, A.; Weisbecker, P.; Geandier, G.; Aeby-Gautier, E.; Appolaire, B. Austenitization and Precipitate Dissolution in High Nitrogen Steels: And In-Situ High Temperature
X-ray Synchrotron Diffraction Analysis Using the Rietveld Method. Mater. Sci. Eng. A 2005, 393, 63–70.
38. Caballero, F. G.; Capdevila, C.; García de Andrés, C. Influence of Scale Parameters of Pearlite on the Kinetics of Anisothermal Pearlite-to-Austenite Transformation in a Eutectoid
Steel. Scr. Mater. 2000, 42, 1159–1165.
39. Cai, X.-L.; Garrat-Reed, A. J.; Owen, W. S. The Development of Some Dual-Phase Steel Structures from Different Starting Microstructures. Metall. Trans. A 1985, 16,
543–557.
40. Datta, D. P.; Gokhale, A. M. Austenitization Kinetics of Pearlite and Ferrite Aggregates in a Low Carbon Steel Containing 0.15 Wt. Pct. C. Metall. Trans. A 1981, 12,
443–450.
41. Durban, C.; Durand, D.; Chevre, P. Determination of Austenitic Transformation during Fast Heat Treatment in Heat Treating. In Proceedings of the 17th ASM Heat Treating
Society Conference Proceedings Including the 1st International Induction Heat Treating Symposium; 1997.
42. Garcia, C. I.; DeArdo, A. J. Formation of Austenite in 1.5 Pct Mn Steels. Metall. Trans. A 1981, 12, 521–530.
43. Garcia, C. I.; DeArdo, A. J. Formation of Austenite in Low Alloy Steels. In International Conference on Solid-Solid Phase Transformations; Aaronson, H. I., et al., Eds.;
TMS-AIME: Warrendale, PA, 1982; p 855.
44. Judd, R. R.; Paxton, H. W. Kinetics of Austenite Formation from a Spheroidized Ferrite–Carbide Aggregate. Trans. Metall. Soc. AIME 1968, 242, 206–215.
45. Klier, E. P.; Troiano, A. R. Ar in Chromium Steels. Trans. Metall. Soc. AIME 1945, 162, 175–185.
46. Lenel, U. R. TTT Curves for the Formation of Austenite. Scr. Metall. 1983, 17, 471–474.
47. Molinder, G. A Quantitative Study of the Formation of Austenite and the Solution of Cementite at Different Austenitizing Temperatures for a 1.27% Carbon Steel. Acta Metall.
1956, 4, 565–571.
48. Nemoto, M. The Formation of Austenite from Mixtures of Ferrite and Cementite as Observed by HVEM. Metall. Trans. A 1977, 8, 431–437.
49. Puskar, J. D.; Dykhuizen, R. C.; Robino, C. V.; Kelley, J. B.; Burnett, M. E. Austenite Formation Kinetics during Rapid Heating in a Microalloyed Steel. In 41st Mechanical
Working and Steel Processing Conference; The Iron and Steel Society (ISS), 1999; Vol. 37, pp 625–635.
50. Roberts, G. A.; Mehl, R. F. The Mechanism and the Rate of Formation of Austenite from Ferrite–Cementite Aggregates. Trans. ASM 1943, 31, 613–650.
51. Roosz, A.; Gacsi, Z.; Fuchs, E. G. Isothermal Formation of Austenite in Eutectoid Plain Carbon Steel. Acta Metall. 1983, 31 (4), 509–517.
52. Schmidt, E.; Damm, E. B.; Sridhar, S. On the Rate and Mechanism of Interface Migration during Austenitization of 4118 Steel. In Proceedings of International Conference on
New Developments in Long and Forged Products: Metallurgy and Applications. 2006; AIST: Warrendale, PA:, 2006; pp 203–215.
53. Schmidt, E.; Damm, E. B.; Sridhar, S. A Study of Diffusion- and Interface-Controlled Migration of the Austenite/Ferrite Front during Austenitiztion of a Case-Hardenable Alloy
Steel. Metall. Trans. A 2007, 38, 244–260.
54. Speich, G. R.; Demarest, V. A.; Miller, R. L. Formation of Austenite during Intercritical Annealing of Dual Phase Steels. Metall. Trans. A 1981, 12, 1419–1428.
55. Speich, G. R.; Szirmae, A. Formation of Austenite from Ferrite and Ferrite–Carbide Aggregates. Trans. Metall. Soc. AIME 1969, 245, 1063–1074.
56. Wycliffe, P.; Purdy, G. R.; Embury, J. D. Austenite Growth in the Intercritical Annealing of Ternary and Quaternary Dual-Phase Steels. In Fundamentals of Dual Phase Steels;
TME-AIME: Warrendale, PA:, 1981; pp 58–83.
57. Wycliffe, P.; Purdy, G. R.; Embury, J. D. Growth of Austenite in the Intercritical Annealing of Fe–C–Mn Dual Phase Steels. Can. Metall. Q. 1981, 20 (3), 339–350.
58. Yang, D. Z.; Brown, E. L.; Matlock, D. K.; Krauss, G. The Formation of Austenite at Low Intercritical Annealing Temperatures in a Normalized 0.08C–1.45Mn–0.21Si Steel.
Metall. Trans. A 1985, 16, 1523–1526.
59. Yang, J. R.; Bhadeshia, H. K. D. H. Reaustenisation in Steel Weld Deposits. In Proceedings of an International Conference on Welding Metallurgy of Structural Steels;
The Metallurgical Society of the AIME: Warrendale, PA, 1987; pp 549–563.
60. Yang, J. R.; Bhadeshia, H. K. D. H. The Bainite to Austenite Transformation or Reaustenitisation from Bainite. In Proceedings of an International Conference: Phase Trans-
formations ’87; Institute of Metals: London, 1988; pp 203–206.
61. Yang, J. R.; Bhadeshia, H. K. D. H. Reaustenitisation Experiments on Some High-strength Steel Weld Deposits. Mater. Sci. Eng. A 1989, 118, 155–170.
62. Yang, J. R.; Bhadeshia, H. K. D. H. Continuous Heating Transformation of Bainite to Austenite. Mater. Sci. Eng. A 1991, 131, 99–113.
63. Epp, J.; Surm, H.; Kessler, O.; Hirsch, T. In-Situ X-ray Investigations and Computer Simulation during Continuous Heating of a Ball Bearing Steel. Metall. Trans. A 2007, 38,
2371–2378.
64. Schmidt, E.; Wang, Y.; Sridhar, S. A Study of Nonisothermal Austenite Formation and Decomposition in Fe–C–Mn Alloys. Metall. Trans. A 2006, 37, 1799–1810.
65. Zhang, X.; Liu, W.; Sun, D.; Li, Y. The Transformation of Carbides during Austenitization and Its Effect on the Wear Resistance of High Speed Steel Rolls. Metall. Trans. A 2007,
38, 499–505.
66. Caballero, F. G.; Capdevila, C.; García de Andrés, C. Influence of Pearlite Morphology and Heating Rate on the Kinetics of Continuously Heated Austenite Formation in
a Eutectoid Steel. Metall. Trans. A 2001, 32, 1283–1291.
67. Caballero, F. G.; Capdevila, C.; García de Andrés, C. Kinetics and Dilatometric Behaviour of Nonisothermal Ferrite – Austenite Transformation. Mater. Sci. Technol. 2001, 17,
1114–1118.
68. Caballero, F. G.; Capdevila, C.; García de Andrés, C. Modeling of Kinetics and Dilatometric Behaviour of Austenite Formation in a Low-Carbon Steel with a Ferritic Plus Pearlite
Initial Microstructure. J. Mater. Sci. 2002, 37, 3533–3540.
69. García de Andrés, C.; Caballero, F. G.; Capdevila, C. Dilatometric Characterization of Pearlite Dissolution in 0.1C–0.5Mn Low Carbon Low Manganese. Scr. Mater. 1998, 38,
1835–1842.
70. García de Andrés, C.; Caballero, F. G.; Capdevila, C.; Bhadeshia, H. K. D. H. Modeling of Kinetics and Dilatometric Behavior of Nonisothermal Pearlite-to-Austenite Trans-
formation in an Eutectoid Steel. Scr. Mater. 1998, 39 (6), 791–796.
71. Kapoor, R.; Kumar, L.; Batra, I. S. A Dilatometric Study of the Continuous Heating Transformations in 18wt.% Ni Maraging Steel Grade 350. Mater. Sci. Eng. A 2003, 352,
318–324.
72. Puskar, J. D.; Dykhuizen, R. C.; Robino, C. V.; Kelley, J. B.; Burnett, M. E. Austenite Formation Kinetics during Rapid Heating in a Microalloyed Steel. Iron Steelmaker 2000, 27,
27–34.
Austenite Formation and Microstructural Control in Low-Alloy Steels 361

73. Reed, R. C.; Akbay, T.; Shen, Z.; Robinson, J. M.; Root, J. H. Determination of Reaustenitisation Kinetics in a Fe–0.4C Steel Using Dilatometry and Neutron Diffraction. Mater.
Sci. Eng. A 1998, 256, 152–165.
74. Grach, G.; Lusk, M.T.; Jou, H.-J.; Ludtka, G.; Elliot, W.H.; Shick, D.; Dicastanzo, G.; Walton, H. Effect of Austenitization Hold Time and Temperature on Matrix Carbon Content
and Martensite Kinetics in 52100. Unpublished draft, 2004.
75. Lyasotsky, I. V.; Shtansky, D. V. Formation of Austenite and the Kinetics of Cementite Dissolution in Steels with a Recrystallized Structure of Granular Pearlite under Laser
Heating. Phys. Met. Metallogr. 1993, 75 (1), 77–82.
76. Miokovic, T.; Schulze, V.; Vöhringer, O.; Löhe, D. Prediction of Phase Transformations during Laser Surface Hardening of AISI 4140 Including the Effects of Inhomogeneous
Austenite Formation. Mater. Sci. Eng. 2006, 435–436, 547–555.
77. Rödel, J.; Spies, H.-J. Calculation of Temperature Fields and Austenite Formation during Electron Beam Hardening. In Surface Modification Technologies VIII: Proceedings of
the Eighth International Conference on Surface Modification Technologies; Nice, France, pp 638–650.
78. Rödel, J.; Spies, H.-J. Modeling of Austenite Formation during Rapid Heating. Surf. Eng. 1996, 12, 313–318.
79. Yakovleva, I. L.; Schastlivtsev, V. M.; Tabatchikova, T. I. Experimental Observation of Diffusionless Formation of Austenite in a Steel with Pearlitic Structure upon Laser Heating.
Phys. Met. Metallogr. 1993, 76 (2), 179–187.
80. Shtansky, D. V.; Inden, G. Phase Transformation in Fe–Mo–C and Fe–W–C Steels – II. Eutectoid Reaction of M23C6 Carbide Decomposition during Austenitization. Acta Mater.
1997, 45, 2879–2895.
81. Babu, S. S.; Elmer, J. W.; Vitek, J. M.; David, S. A. Time-Resolved X-ray Diffraction Investigation of Primary Weld Solidification in Fe–C–Al–Mn Steel Welds. Acta Mater. 2002,
50, 4763–4781.
82. Elmer, J. W.; Palmer, T. A.; Babu, S. S.; Zhang, W.; DebRoy, T. Direct Observations of Austenite, Bainite, and Martensite Formation during Arc-Welding of 1045 Steel Using
Time-Resolved X-ray Diffraction. Weld. J. 2004, 244S–253S.
83. Palmer, T. A.; Elmer, J. W.; Babu, S. S. Observations of Ferrite/Austenite Transformations in the Heat Affected Zone of 2205 Duplex Stainless Steel Spot Welds Using Time
Resolved X-ray Diffraction. Mater. Sci. Eng. A 2004, A374, 307–321.
84. Zhang, W.; Elmer, J. W.; DebRoy, T. Kinetics of Ferrite to Austenite Transformation during Welding of 1005 Steel. Scr. Mater. 2002, 46, 753–757.
85. Caballero, F. G.; Capdevila, C.; Garcia de Andres, C. Modeling of the Interlamellar Spacing of Isothermally Formed Pearlite in a Eutectoid Steel. Scr. Mater. 2000, 42, 537–542.
86. Dykhuizen, R. C.; Robino, C. V.; Knorovsky, G. A. A Method for Extracting Phase Change Kinetics from Dilatation for Multistep Transformations: Austenitization of a Low Carbon
Steel. Metall. Trans. B 1999, 30, 107–117.
87. Katsamas, A. I. A Computational Study of Austentite Formation Kinetics in Rapidly Heated Steels. Surf. Coat. Technol. 2007, 201, 6414–6422.
88. Shtansky, D. V.; Nakai, K.; Ohmori, Y. Pearlite to Austenite Transformation in an Fe–2.6Cr–1C Alloy. Acta Mater. 1999, 47 (9), 2619–2632.
89. Zhao, L.; Vermolen, F. J.; Wauthier, A.; Sietsma, J. Cementite Dissolution at 860  C in an Fe–Cr–C Steel. Metall. Trans. A 2006, 37, 1841–1850.
90. Liu, Z. K.; Höglund, L.; Jönsson, B.; Ågren, J. An Experimental and Theoretical Study of Cementite Dissolution in an Fe–Cr–C Alloy. Metall. Trans. A 1991, 22, 1745–1752.
91. Liu, Z. K.; Ågren, J. Morphology of Cementite Decomposition in an Fe–Cr–C Alloy. Metall. Trans. A 1991, 22, 1753–1759.
92. Li, Z.-D.; Miyamoto, G.; Yang, Z.-G.; Furuhara, T. Kinetics of Reverse Transformation from Pearlite to Austenite in an Fe-0.6 Mass Pct C Alloy and the Effects of Alloying
Elements. Metall. Mater. Trans. A 2011, 42, 1586–1596.
93. Chae, J.-Y.; Jang, J.-H.; Zhang, G.; Kim, K.-H.; Lee, S. L.; Bhadeshia, H. K. D. H.; Suh, D.-W. Dilatometric Analysis of Cementite Dissolution in Hypereutectoid Steels Containing
Cr. Scr. Mater. 2011, 65, 245–248.
94. Miyamoto, G.; Usuki, H.; Li, Z.-D.; Furuhara, T. Effects of Mn, Si, and Cr Addition on Reverse Transformation at 1073 K from Spheroidized Cementite Structure in Fe-0.6 Mass%
C Alloy. Acta Mater. 2010, 58, 4492–4502.
95. Palizdar, Y.; San Martin, D.; Brown, A. P.; Ward, M.; Cochrane, R. C.; Brydson, R.; Scott, A. J. Demonstration of Elemental Partitioning during Austenite Formation in Low-
Carbon Aluminium Alloyed Steel. J. Mater. Sci. 2011, 46, 2384–2387.
96. Lee, S. J.; Mola, J.; De Cooman, B. C. Conversion Model for the Martensitic Transformation of Banded Austenite in a Ferrite Matrix. Metall. Mater. Trans. A 2012, 43,
4921–4925.
97. Azizi-Alizamini, H.; Militzer, M.; Poole, W. J. Austenite Formation in Plain Low-Carbon Steels. Metall. Mater. Trans. A 2011, 42, 1544–1557.
98. Kulakov, M.; Poole, W.J.; Militzer, M. The Effect of the Initial Microstructure on Recrystallization and Austenite Formation in DP600 Steel. Submitted to MMTA December 2012.
99. Rudinzki, J.; Bottger, B.; Prahl, U.; Bleck, W. Phase-field Modeling of Austenite Formation from a Ferrite Plus Pearlite Microstructure during Annealing of Cold-Rolled Dual-Phase
Steel. Metall. Mater. Trans. 2011, 42, 2516–2525.

You might also like