PHYSICAL METALLURGY OF THE THERMOMECHANICAL PROCESSING
OF MICROALLOYED STEELS
A. J. DeArdo
Basic Metals Processing Research Institute
Department of Materials Science and Engineering
University of Pittsburgh, Pittsburgh, PA 15261, USA
ABSTRACT
The fundamental metallurgy of thermomechanical processing (TMP) has become much better
understood over the last decade. An appreciation of the driving forces and retarding forces has led
to both a better understanding of existing practices and the development of new practices.
Furthermore, the distinction between plate and strip rolling, for example, has become better
defined, quantified and understood. This understanding has led to improved products, increased
productivity and better consistency. The information generated, first in plate and then in strip
rolling, is slowly being adopted in the structural rolling and forging industries. Finally, the role of
TMP has been extended, with the new ultra-low carbon steels, to include the control of
stabilization reactions and texture development during rolling for downstream benefit in the cold
mill. This paper reviews the basic physical metallurgy of TMP, discusses various aspects of alloy
design, and presents illustrations of modern views of plate processing.
Keywords: hot rolling, plate steel, recrystallization, strip steel, ultra-low carbon steel
INTRODUCTION
The addition of elements such as niobium, titanium and vanadium to the early pearlite-reduced
steels was used to facilitate the austenite conditioning process (1-19) as well as to lower the
transformation temperature (20-22) and to possibly provide precipitation hardening of the ferrite
(23). Exhaustive research has shown that the effectiveness of austenite conditioning can be
quantified through the use of the parameter SV, the total interfacial area of near planar crystalline
defects such as grain boundaries, deformation bands and incoherent twin boundaries. Since it is
these defects that act as nucleation sites for ferrite during transformation, larger values of SV were
found to be associated with finer ferrite grains, Figure 1. It is now well understood that fine ferrite
grains are a result of high SV and low Ar3 temperatures (17-18,24-31).
When is isothermally hot deformed and held to simulate passes in an industrial practice such
as one of those described above, and then quenched to RT to observe the resulting microstructure,
the pattern of behavior which is found is shown schematically in Figure 2.
When the deformation occurs above T95, a recrystallized grain structure results. However,
when the pass strain occurs below T5, a pancaked microstructure is observed. Passes which take
place between these two temperatures lead to detrimental mixed grain structures.
It is apparent from the phenomenology shown in Figure 3 (24) and from the quantification of
such data that
Sv = 2/D,
for T>TRXN
(1)
for T<TRXN
(2)
and
Sv = (1 + R + 1/R)/D + 0.63(%R - 30)
where R is the average aspect ratio of deformed austenite grains (32), that large values of SV can
be achieved through: (i) refining the starting grain size, (ii) increasing the grain boundary area
per unit volume by changing the grain shape, e.g., pancaking, and (iii) increasing the density of
intragranular crystalline defects such as deformation bands and/or incoherent twin boundaries.
Hence, large values of Sv can be achieved through grain refinement, grain pancaking, or a
combination of the two. The first contribution can occur only through control of the reheating
practice and by applying most of the passes above the temperature T95 described in Figure 2. The
second contribution can occur only through applying most of the passes below T5, described in
Figure 2. In this second case, the more total deformation applied below T5, the higher would be
the Sv caused by this approach, i.e., pancaking.
Two comments can be made regarding the general behavior of shown in Figure 2. The first is
that the diagram will vary with the type of process being considered, even for the same steel and
reheating conditions. This is a result of the difference in mill characteristics shown in Table I and
in the response of the material to these differences. The second is that the overall behavior shown
in Figure 2 represents the result of the competition between the driving forces trying to eliminate
the excess crystalline defects induced during the processing and the retarding forces that slow or
retard this change. Obviously, when the driving force exceeds the pinning force and sufficient
diffusivity of pertinent atomic species exists, the movement of defects leading to their reduction or
elimination can proceed. The migration of dislocations and low and high angle boundaries
resulting in lower numerical and energetic densities are, in fact, what is meant by recovery,
recrystallization and grain coarsening. However, when the retarding force exceeds the driving
force, or when insufficient atomic mobility exists, then the structural restoration processes will not
occur in the time available. For static restoration, which is the predominant mode possible during
industrial processing, the time available is the interpass time shown in Table I.
The purpose of this paper is to review the physical metallurgy that underlies the reactions that
occur during the hot rolling of steel: grain coarsening, recovery and recrystallization. Particular
emphasis will be placed on the forces which drive and retard each type of reaction, and the
conditions under which each reaction may or may not occur. It is hoped that a better
understanding of these forces can lead to further improvements in the alloy and process design of
steels for the future.
BACKGROUND PHYSICAL METALLURGY
Grain Coarsening The coarsening of austenite grains has been treated by several investigators
(33-35). The approach adopted here is that of Gladman who treated the onset of abnormal grain
coarsening or secondary recrystallization (36-38). It is often the abnormal rather than the normal
grain coarsening which is of greatest importance in TMP. Gladman suggested that grain
coarsening is accompanied by two types of boundary motion: the inward motion of boundaries of
shrinking grains and the outward motion of boundaries of growing grains (38). The first event is
accompanied by an energy decrease (driving force) given by 3/2Ro, while the second is
accompanied by an energy increase (retarding force) given by 2/R, where is the grain boundary
energy, Ro is the average radius of a matrix of tetrakaidecaherdral grains, and R is the radius of the
growing grain. The net driving force is then [3/2Ro - 2/R]. An energy balance indicated that
only grains with a size larger than 4/3 that of the average could grow. This clearly shows the
importance of homogeneity in the grain size distribution. According to this view, very narrow
grain size distributions would be expected to exhibit little or no grain coarsening, independent of
average grain size. This clearly shows the value of determining the complete grain size histogram
or size distribution rather than just an average grain size since the width of the distribution will
indicate the coarsening tendency.
Coarsening Inhibition Using Particles While the tendency for grain coarsening will always be
present in commercial steels, it is important to realize that this coarsening tendency can be slowed
or eliminated by the presence of suitable forces that can inhibit boundary motion. Two of the more
important mechanisms are particle pinning and solute drag. When a boundary intersects a particle,
a portion of the boundary is eliminated. To pull the boundary away from the particle requires the
creation of additional boundary area, and, hence, the expenditure of energy or equivalently the
application of a force. This force is the force for grain coarsening. Gladman's treatment of grain
coarsening also included the effects of particles (36-38). Gladman's analysis is based on the earlier
of work of Zener (39). Gladman's energy balance considering the shrinking of small grains, the
expanding of large grains, and the pulling of boundaries away from particles led to what is
considered the Gladman equation
rC = 6 RO f [ (3/2 - 2/Z)]-1
(3)
where rC is the critical or maximum particle size required to suppress coarsening, RO is the radius
of the grain of average size, f is the volume fraction of particles, and Z is the ratio of radii R/RO of
growing to average grains. Hence, for any average grain size, there exists a combination of
particle size and particle volume fraction that can suppress coarsening. The Gladman equation is
shown in Figure 4, assuming Z = 1.5 (38).
The negatively sloping lines represent combinations of rC and f that can pin all grain sizes which
fall above the line. The finest grain sizes that can be stabilized by particles in commercial steels are
shown in the lower hatched area. The upper hatched area shows the finest grains that can be
anchored by large inclusions. The nearly two orders of magnitude difference in grain sizes that can
be stabilized by these two very different particle distributions is striking.
Coarsening Inhibition Using Solute Drag When solute atoms are located on and attached to
a grain boundary, the migration rate of the boundary resulting from a given driving force will be
influenced by the solute atoms. The theory of solute drag has been the subject of numerous
studies and reviews (40-42). Most of this work was conducted in recrystallization experiments
where, as discussed above, the driving forces are rather high. Grain coarsening, on the other hand,
takes place at very low driving forces.
According to Cahn's theory (41), the rate of grain growth or coarsening in the presence of solute,
G, can be given as
G2 = ( 2 VM n / t) / (' + C)
(4)
where is the specific grain boundary energy per unit volume, VM is the molar volume of
austenite, n is the isothermal grain coarsening law exponent, t is the coarsening time, ' is the
reciprocal boundary mobility of "pure" austenite, is the reciprocal boundary mobility at unit
concentration solute, and C is the bulk solute concentration. The linear relationship predicted by
this theory between the G-2 -vs- C has, in fact, been observed (42). The coarsening rate will fall
with increasing solute levels. The choice of solute will influence the values of , n, VM , and ;
hence, each solute would be expected to exert its own particular influence on G.
The combined Cahn-Lcke-Stwe (CLS) theory of solute-altered boundary motion is shown in
equation (5) for the conditions of both high and low driving force. (41,43,44) The driving
pressure P for grain coarsening is given by
P = V/M + COV/(1+'V2)
(5)
where V and M are the boundary velocity and mobility, respectively, CO is equilibrium solute
level, and and ' are constants depending upon the model used. The first term represents the
high driving force regime where the boundary motion is taken to be only weakly slowed by solute.
The second term represents the low driving force regime where the boundary mobility is strongly
altered (slowed) by solute. One of the main weaknesses of applying these theories is the
estimation of the actual grain boundary solute concentration. Perhaps an improvement will be
possible when more direct measurement of these concentrations become available. The Atom
Probe Field Ion Microscope (APFIM) is particularly well-suited in this regards since the mass
spectrometer (AP) defines the atomic species and the microscope function determines the solute
distribution, e.g., mono-versus multi-layer coverage, on the boundary (45).
Recrystallization Driving Force The driving force for static recrystallization can be taken to
be the stored energy of deformation or the difference in dislocation density between the deformed
or recovered matrix and the recrystallized matrix. This can be estimated by using the approach of
Keh and Weissmann (43) who related the increase in flow stress caused by work hardening to the
increase in dislocation density. Their expression is
= O + b
(6)
where is the observed increase in flow stress caused by work hardening, O is the initial yield
strength, is a constant, is the shear modulus, b is the Burgers vector and is the change in
dislocation density. The driving force for recrystallization is then given by
FRXN = b2 / 2
(7)
As mentioned above, driving forces resulting from individual hot roll pass simulations have been
found to be on the order of 20 MN/m2 using this technique (46-48). Of course, it is possible to
directly estimate the dislocation density in metals and alloys that do not undergo a phase
transformation during cooling through the use of transmission electron microscopy. The stored
energy of deformation is known to be influenced by starting condition or microstructure, strain
magnitude, deformation temperature, and strain rate (1-5).
Pinning Forces Resisting Recrystallization To retard recrystallization, the pinning forces
developed by the strain-induced precipitate must be larger than the 20 MN/m2 driving force
available to cause recrystallization. The pinning force can be expressed as
(8)
FPIN = 4 r NS
where r is the particle radius, is the interfacial energy of the austenite grain boundary, and NS is
the number of particles per unit area. Three models have been proposed for NS to calculate the
pinning forces exerted by arrays of microalloy carbonitride precipitates (49,50). These are the
rigid boundary, the flexible boundary and the subgrain boundary models. Recent work has shown
that the subgrain model of Hansen et al.(50) is probably the most realistic (46-48). This model
predicts the pinning force to be
FSPIN = 3 fV l [ 2 r2 ] -1
(9)
where is the austenite grain boundary energy, fV is the volume fraction particles, l is the subgrain
size, and r is the particle radius. From equation 9 it appears to be rather simple to measure an
average r and to measure or calculate the fV of the particles. An example of the precipitates
causing this pinning effect in a Nb steel is shown in Figure 5.
It is immediately apparent from Figure 5 that the particles are not randomly distributed, but
appear to have formed on the subgrain boundaries of deformed and recovered austenite. Clearly,
an average value for NS would be inappropriate since the distribution of particles is so localized. It
has been found that true localized pinning forces at the sites for preferential precipitation are far
higher than that expected from a uniform distribution of particles (46-48). Equation 9 shows that
high pinning forces will be promoted by high values of fV and small particle sizes. Those locations
with very high localized values of fV, , i.e. grain and subgrain boundaries as shown in Figure 5, will
have especially high magnitudes of FPIN.
In order to have a sufficient number of passes below T5 in Figure 2, it is clearly necessary to
have T5 in the correct temperature range for optimized controlled rolling. This means that the
strain-induced precipitation should begin at a temperature which is well up in the range of
temperatures where the rolling passes actually occur. If the precipitate forms at too high of a
temperature it will not be able to generate a sufficiently high pinning force to suppress static
recrystallization during rolling at lower temperatures. If it forms at too low of a temperature, T5
will be too close to the finishing temperature to allow enough passes to occur thereby limiting the
achievable SV.
It appears that there are three conditions required to have particle pinning forces large enough
to suppress recrystallization. First, a large supersaturation is needed in the temperature range that
is proper from the standpoint of both the solubility product and the rolling schedule. Second, a
conditioned microstructure containing a large number of sites for precipitation is needed. Third,
sufficient time is required to nucleate and grow a large number, i.e., a large fV, of small
precipitates.
Unless all three of these conditions are met, it would be impossible to generate the particle
pinning forces necessary to suppress static recrystallization. Plate mill rolling on a reversing mill
or rolling of heavy sections are probably the most likely scenario where recrystallization
suppression through particle pinning can occur. The interpass times and total mill processing
times from 1000C to 800C are probably much too short for this pinning in strip and bar mills.
Pinning Effects Caused by Solutes There is no adequate theoretical treatment of the
suppression of static recrystallization during multi-pass hot deformation by solute atoms. At this
point in time, there are treatments of solutes on high angle boundary mobility and of interactions
between solutes and dislocations. The influence of solutes on grain boundary motion was
discussed above in equation 5. The theories of solute- dislocation interactions form the basis for
solid solution hardening, but how this pertains to subgrain formation and the nucleation of
recrystallization is unclear.
The intensity with which different solute additions act to suppress static recrystallization has
been measured and is shown in Table II (51). The SRP values shown in Table II indicate the ratios
of time for a constant degree of fractional softening for the microalloyed steel as compared to an
equivalent carbon steel. The extreme effectiveness of Nb in solution is obvious.
Summary Thirty years of research has revealed that the microstructure of hot rolled austenite
can be rationalized by considering both the steel and its processing. One of the fundamental issues
is whether or not static recovery, recrystallization and abnormal grain coarsening can occur in the
available interpass time. Another issue is how additions of microalloying elements might thwart
these tendencies. Pinning forces based on particles and/or solutes are tools that are available to
control the evolution of microstructure of hot rolled austenite.
OPTIMIZATION OF COMPOSITION FOR PLATE ROLLING
Plate rolling often involves large widths which require a large number of relatively light passes
because of mill limitations that restrict the separating force and torque in any given pass. Also,
plate rolling often occurs on reversing mills, where the interpass time can be on the order of 10 to
30 seconds. Since strain-induced precipitation can occur within a few seconds (46,47), retardation
of recrystallization using particles is an option in plate rolling. Retardation of recrystallization
using solutes alone is not likely because of their small pinning forces and the long
times available between passes. Solutes, however, are quite adequate in suppressing grain
coarsening.
Alloy Design Considerations for Controlled Rolling of Plate What follows is a discussion,
much of which has been published elsewhere (17,19,26,28,29,52), of how microalloying additions
can be used to tailor a steel to a given TMP deformation practice. In this discussion, straininduced precipitation is known to form on austenite grain and subgrain boundaries causing the
retardation of recrystallization. The precipitation behavior in austenite of the various microalloy
carbides, carbonitrides and nitrides can be best understood through the use of the solubility
equations. Figure 6 illustrates the behavior of several popular systems (52).
Each system has been assumed to be reheated and then cooled to different temperatures. The
supersaturation of the microalloying element at each temperature was calculated using standard
solubility products. The supersaturation was determined by comparing the amount of
microalloying element dissolved during reheating to the amount in solution at each temperature.
Precipitation will occur when two criteria are met: (i) the supersaturation reaches a critical value
and (ii) there is adequate atomic mobility of the metallic element. The curves of Figure 6 show
that TiN has the lowest and VC has the highest solubility in austenite.
Perhaps the most important concept governing the choice of microalloying additions for
response to low temperature controlled rolling is the T5 temperature shown in Figure 2. One of
the most interesting, powerful and important roles played by microalloying additions is their
influence on T5 or the recrystallization-stop temperature. This effect is shown in Figure 7 (53).
Although all of the microalloying additions raise the T5, the influence of Nb is especially
pronounced. The trend shown in Figure 7 can be understood with the help of Figure 6 which
shows that TiN precipitates at temperatures that are too high and VN and VC at temperatures that
are too low to be effective in forming strain-induced precipitates that can retard static
recrystallization over the ideal temperature range. It is important to note that the very low T5 of
about 800C for the C- Mn- Si base steel is below the finishing temperature of all but the most
modern and sophisticated plate mills. This means that it is nearly impossible to control roll nonmicroalloyed steels.
The mechanism by which microalloying additions act to raise T5 has been the subject of
numerous investigations over the last quarter century. Mechanisms based on solute drag (54-56)
and particle pinning have been proposed (57-59). It is now fairly widely accepted that it is the
strain-induced precipitates that suppress the nucleation of statically recrystallized grains during the
interpass time following a hot deformation pass (46,48-50,60). For this mechanism to hold, the
pinning forces developed by the particle dispersions must: (i) be larger than the driving force for
recrystallization; (ii) must be found superior in the temperature range that corresponds to the T5,
which itself must be determined by independent means; and (iii) must be less than the driving force
at temperatures above T95 , which itself must be determined by independent means.
Recent work has shown that all of these criteria were met in the microalloyed steels investigated
(46,48). In this study, the recrystallization-stop or T5 temperatures were determined for a series of
low carbon steels containing 0.08C- 1.5 Mn- 0.3Si and varying amounts of Nb. The straininduced Nb(CN) precipitation was observed for each steel at 10C below its respective T5
temperature. Pinning forces were calculated from these observations for the precipitation in the
austenite grain interiors or matrix (M) and in the vicinity of the austenite grain boundaries (B).
Pinning forces were also calculated using solubility relations which would yield some average
value (EQ). The results are shown in Figure 8 for steels E3, containing 0.02 Nb and exhibiting a
T5 of 940C, and E4, containing 0.09 Nb and exhibiting a T5 of 1030C (47,48).
The pinning forces in Figure 8 were calculated at 1100C, 900C, and at 10C below the T5 for
each steel. It is quite clear that the largest pinning force occurs near the grain boundaries.
Considering that the driving force for recrystallization is about 22 MPa, it appears that the
precipitation in the vicinity of grain boundaries is responsible for the suppression of nucleation of
recrystallization. This is not especially surprising since the grain boundary regions are the most
preferred sites for all heterogeneously nucleated reactions, i.e. precipitation, recrystallization, and
nucleation of ferrite. The comparison between driving and pinning forces found in this experiment
is revealed that precipitation on grain boundaries appeared to control T5 or TRXN (47,48).
When austenite is deformed below its T5, an essentially cold worked or recovered
microstructure results. This type of microstructure is shown in Figure 9 for a partially
transformed, controlled rolled austenite in a Nb-B steel (52). Note how the proeutectoid ferrite
has decorated the crystalline defects that comprise the SV for this deformed austenite. The role of
SV in nucleating ferrite is clear in this figure.
Alloy Design for Recrystallization Controlled Rolling The CCR approach to austenite
conditioning is based on having a high T5, which requires that the bulk microalloying composition
of the steel and the reheating temperature combine to provide an adequate supersaturation for
strain-induced precipitation which must occur at a temperature rather early in the rolling process.
This means that the rather large pinning force required to retard static recrystallization must be
developed during the rolling process, and is not present after reheating. The RCR approach to
austenite conditioning, on the other hand, is based on having a grain coarsening inhibition system
and a low T5. The first requirement means that a pinning force must be present immediately after
reheating. The magnitude of this pinning force must be small enough to permit static
recrystallization of austenite to occur, but large enough to suppress grain coarsening. The second
requirement means that strain induced precipitation must be minimized or avoided. Clearly, the
alloy design required for RCR processing is rather restrictive, and most steels will not be amenable
to this type of processing.
The concept of RCR processing is shown schematically in Fig. 10, which compares
conventional hot rolling (CHR) and RCR. The benefits of grain refinement through
recrystallization are negated in the CHR practice because of the rapid grain coarsening that follows
static recrystallization. However, if a pinning force of appropriate magnitude is present, then the
grain coarsening would be suppressed and a fine grain size would result.
One approach to alloy design for RCR processing is based on TiN as the grain coarsening
inhibition system and V(CN) as the ferrite strengthening system (61-63). The TiN which
precipitates in the solid generates pinning forces of the ideal magnitude, i.e. large enough to
suppress grain coarsening of austenite grains recrystallized after hot deformation, Figure 11 , but
small enough to permit the static recrystallization to occur (64,65).
The V(CN) is also ideal since little strain-induced precipitation would be expected at the high
finishing temperatures, approximately 950C, used in this practice. The V-Ti-N microalloyed
steels respond to RCR processing very well, both in terms of austenite conditioning (62,66) and in
terms of final properties (66). The yield strengths in these steels are limited to about 500 MPa
because of some loss in precipitation hardening caused by the epitaxial precipitation of a portion of
the V(CN) on the pre-existing TiN (67). The transition temperatures in these steels were
excellent ranging from -70 to 130C, depending on strength level. The strength and toughness
of these very fine grained ferrite-pearlite steels are excellent. Other aspects of processing and
properties of similar steels have also been investigated (46,61,63,67).
A second approach to the alloy design for RCR processing, based on the Nb-Mo system, has
recently been developed (45). The inhibition of grain coarsening using this system is also
illustrated in Figure 11. Recent work has shown that the pinning force is caused mainly by the
combination of the presence of fine Nb(CN) which remained undissolved during reheating and the
segregation of Mo at the boundaries. This segregation has been examined in these steels using the
Atom Probe Field Ion Microscope (45). Both C and Mo are shown segregated to the austenite
grain boundaries. The pinning force caused by the Nb(CN) precipitate was determined to be about
86 kN/m2 and that by the Mo was 40 kN/m2. The total was about equal to the estimated driving
force for coarsening.
For RCR processing to work, it is important to know where post-deformation static
recrystallization will occur. Recrystallization maps, determined for the appropriate conditions, are
helpful in this regard. (64,68).
OPTIMIZATION OF COMPOSITION FOR STRIP ROLLING
Rolling on a hot strip mill is quite different than on a plate mill. The widths are narrower, the
rolling speeds are higher, the reductions are greater and the total rolling times are much shorter.
In a typical strip mill the time from slab drop out to exiting the last finishing pass is only about one
to two minutes. Comparable times are five to ten minutes for plate or structural rolling. Since
nearly all of this time is interpass time, static hardening and softening processes are expected to be
much more important and variable in plate and structural rolling than in strip rolling. These
concepts, applied to strip rolling, have been reviewed recently (69).
SUMMARY AND CONCLUSIONS
Quantitative metallography has been a very helpful tool in gaining a better understanding of TMP.
It is now possible to measure both the driving and retarding forces that play an intimate role in
TMP. This quantitative information is permitting a better understanding of current steels and
TMP practices and will hopefully enable improvements to continue to be made in both in the
future.
Plate, strip and sheet steels are obviously very different materials, with very different functional
requirements. It is vitally important to recognize the different processing paths experienced by
these different products. Matching the alloy design with the processing enables both the
processing and final product to be optimized.
REFERENCES
1. Proc. Microalloying 75 (Washington, DC), Union Carbide Corp., New York, 1977.
2. Proc. Hot Deformation of Austenite (Cincinnati), TMS-AIME, New York, 1976.
3. Proc. HSLA Steels: Tech. and Applications (Philadelphia), ASM, Metals Park, OH, 1984.
4. Niobium (San Francisco), TMS-AIME, Warrendale, PA, 1984.
5. Proc. Thermomechanical Processing of Microalloyed Austenite (Pittsburgh), TMS-AIME,
Warrendale, PA, 1984.
6. Proc. High-Strength Low Alloy Steels (Wollongong, NSW), South Coast Printers, Port
Kembla, NSW, 1986.
7. Proc. Accelerated Cooling of Steel (Pittsburgh), TMS-AIME, Warrendale, PA, 1986.
8. Proc. HSLA Steels (Beijing), TMS-AIME, Warrendale, PA, 1986.
9. Proc. Microalloyed Forging Steels (Golden), TMS-AIME, Warrendale, PA, 1987.
10. Proc. Accelerated of Rolled Steel (Winnipeg), Pergamon Press, New York, 1988.
11. Proc. Processing, Microstructure and Properties of HSLA Steels (Pittsburgh), TMS-AIME,
Warrendale, PA, 1988.
12. Proc. Microalloyed HSLA Steels (Chicago), ASM, Metals Park, OH, 1988.
13. Proc. HSLA Steels (Beijing), TMS-AIME, Warrendale, PA, 1992.
14. Proc. Processing, Microstructure and Properties of Microalloyed and Other Modern HSLA
Steels (Pittsburgh), ISS-AIME, Warrendale, PA, 1992.
15. Proc. Low Carbon Steels for the Nineties (Pittsburgh), TMS-AIME, Warrendale, PA, 1993.
16. M. E. Fine, Phase Transformations in Condensed Systems, McMillan, New York, 1964.
17. D. J. Swinden and J. H. Woodhead, J. Iron Steel Inst., 209, (1971), 883.
18. G. R. Speich et al., Proc. Phase Transformations in Ferrous Alloys (Philadelphia), (TMSAIME, Warrendale, PA, 1984), 341-390.
19. A. J. DeArdo, Ref. 6, 20.
20. N. S. Pottore et al., Metall. Trans. A, 22A, (1991), 1871.
21. C. I. Garcia and A. J. DeArdo, Metall. Trans., 12A, (1981), 521.
22. C. I. Garcia and A. J. DeArdo, Proc. Solid-Solid PhaseTransformations, (TMS-AIME,
Warrendale, PA, 1982), 855.
23. D. Webster and J. A. Woodhead, J. Iron and Steel Inst., 202, (1964), 987.
24. I. Kozasu et al., Ref. 1, 120-135.
25. J. D. Grozier, Ref. 1, 241-250.
26. K. Amano et al., Ref. 7, 349-365.
27. A. J. DeArdo, Ref. 7, 97.
28. T. Tanaka, Ref. 10, 187-208.
29. A. J. DeArdo, Can. Met Q., 27, (1988), 141.
30. E. E. Underwood, Quantitative Metallography, (McGraw-Hill, New York, 1968), 77.
31. L. J. Cuddy, Metall. Trans. A, 15A, (1984), 87-98.
32. E. J. Palmiere, University of Pittsburgh, Unpublished Research, 1990.
33. O. O. Miller, Trans. ASM, 43, (1951), 260-289.
34. T. Gladman and F. B. Pickering, J. Iron Steel Inst., 205, (1967), 653-664.
35. N. E. Hannerz and F. DeKazinczy, J. Iron Steel Inst., 208, (1970), 475-481.
36. T. Gladman, Proc. Roy. Soc., 294A, (1966), 298-309.
37. T. Gladman, Proc. Recrystallization and Grain Growth of Multi-Phase and Particle Containing
Materials (Roskilde), (RISO National Laboratories, 1980), 183-192.
38. T. Gladman, Ref. 13, pp. 3-14.
39. C. Zener, Private Communication to C. S. Smith, T.A.I.M.M.E., 175, (1949), 15.
40. K. Lucke and K. Detert, Acta Met., 5, (1957), 628.
41. J. W. Cahn, Acta Met., 13, (1962), 789.
42. P. Gordon and R. A. Vandermeer, Proc. Recrystallization, Grain Growth and Textures, (ASM,
Metals Park, OH, 1966), 205.
43. K. Lucke and H. P. Stuwe, Recovery and Recrystallization in Metals, ed. L. Himmel,
(Interscience Publ., 1963), 171.
44. F. J. Humphreys and M. Hatherly, Recrystallization and Related Annealing Phenomena,
(Pergamon Press, 1996), 114.
45. P. Pouly, University of Pittsburgh, Unpublished Research, 1993.
46. O. Kwon and A. J. DeArdo, Acta Met., 39, (1991), 529.
47. E. J. Palmiere, Ph. D. Thesis, University of Pittsburgh, 1991.
48. E. J. Palmiere, C. I. Garcia, and A. J. DeArdo, Metall. Mater. Trans. A, 25A, (1994), 277-286,
27A, (1996), 951-960.
49. L. J. Cuddy, Ref. 5, 129-140.
50. S. S. Hansen et al., Metall. Trans. A, 11A, (1980), 387.
51. M.G. Akben et al., Acta Met, 31, (1983), 161-174.
52. A. J. DeArdo, University of Pittsburgh, Unpublished Research, 1985.
53. L. J. Cuddy, Plastic Deformation of Metals, (Academic Press, New York, 1975), 129-140.
54. J. J. Jonas and I Weiss, Metals Sci., 13, (1979), 238.
55. M. G. Akben et al., Acta Met., 29, (1981), 111.
56. M. J. Luton, Metall. Trans. A, 11A, (1980), 411.
57. J. D. Jones and A. B. Rothwell, Proc. Deformation Under Hot Working Conditions, (ISI
Publication 108, ISI, London, 1968), 78.
58. A. T. Davenport et al., J. Metals, 27, (1975), 21.
59. A. T. Davenport et al., Ref. 2, 186.
60. A. J. DeArdo et al., ASM Metals Handbook , (ASM International, Metals Park, OH, 1991),
Vol 4, 237.
61. W. Roberts, Ref. 3, 33.
62. Y. Z. Zheng et al., Ref. 3, 85-94.
63. T. Siwecki et al., Ref. 3, 619-634.
64. C. I. Garcia et al., Ref. 14, 395-400.
65. C. I. Garcia et al., Soc. Manufacturing Engineers, Paper MF 92-139, (1992).
66. R. M. Fix et al., Ref. 8, 219-228.
67. S. Zajac et al., Ref. 14, 511-526.
68. C. I. Garcia et al., SAE International Congress, Paper 910143, (1991).
69. A. J. DeArdo, Materials Science Forum, Vols. 284-286, (1998), pp15-26.
Table I Typical Characteristics of Hot
Rolling Processes
Mill
Plate
Reduction, %
15
Strain Rate,
20
Sec-1
Interpass
20
Time, Sec.
Finish T, C 750-950
F
1382-1742
No. Passes
15
Finishing
Stands of
Strip Mill
50
100
25
100
900
1652
6
1000
1832
8
Bar
SRP
(0.1at%)
12
33
70
325
Initial
Suppression
of RXN
T 95%
Partial
RXN
Table II Solute Retardation Parameters
for Static Recrystallization
Element
(0.1%)
V
Mo
Ti
Nb
Complete
RXN
T 5%
Complete
Suppression of
RXN
No RXN
pass
STRAIN
Figure 2 - Schematic illustration of austenite
microstructures resulting from various
deformation conditions (41). T95 and T5 are
the temperature for 95% and 5%
recrystallization, respectively.
SRP
(0.1wt%)
13
20
83
222
Figure 1 - Ferrite grain sizes produced from
recrystallized and unrecrystallized austenite at
various Sv values (31).
Figure 3 - Variation of effective interfacial area
(SV) with rolling below the recrystallization
temperature for the 0.03% Nb steel. (24)
DEFORMATION TEMP RANGE
1000
INCLUSIONS
Particle
Size
TiN
80
.02 Ti- .01 N
60
r = 1000 nm
100
100
r = 10000 nm
.2 Ti- .01 N
NbC
VN
VC
40
r = 100 nm
10
20
GRAIN
REFINEMENT
.04 Nb-.1 C
1300 1200
.12 V - .01 N
1100 1000 900
800
.12 V - .1 C
700
TEMPERATURE, C
Figure 6 - Precipitation potential of various
microalloying systems (60).
r = 10 nm
1
0.01
0.10
1.00
% By Volume of Particles
Figure 4 - Effect of particles on inhibition of
grain growth (38).
Nb
Ti
Al
Figure 5 - Dark field TEM micrograph of strain
induced precipitation of Nb(CN) in a 0.09 wt%
C-0.07Nb steel. Specimen reheated to 1250C
and air cooled to room temperature.
Figure 7. The increase in recrystallization
temperature with the level of microalloy
solutes in a 0.07C, 1.40Mn, 0.25Si steel (53).
CHR
D
Fpin
RCR
PN
PN + 1
PN + 2
Figure 10 - Grain size evolution through
succeeding passes depending on TMP.
Figure 8 - Calculated Nb(CN) pinning force for
steel E4 and E3, at austenite grain boundaries
(B), within austenite grains (M) and what
would be predicted from solubility products
(EQ), based on subgrain boundary model
(47,48).
100
80
C - Mn
60
40
V - Ti - N
20
BHS - 1
0
20
40
60
80
100
Time (sec)
Figure 11 - Grain coarsening behavior of
austenite after deformation (64,65).
Figure 9 - Dark field optical micrograph of
nucleation of ferrite at deformed austenite grain
boundaries, deformation bands (A), and
annealing twins (B) (52).