The Influence of Niobium Microalloying on Austenite Grain ...

7 downloads 70286 Views 2MB Size Report
automotive applications with a goal of increasing carburiz- ... 1) Department of Mechanical Technology, Buraydah College of Technology, Buraydah, Saudi Arabia. E-mail: ... Colorado School of Mines, Golden, CO 80401. E-mail: ...
ISIJ International, Vol. 47 (2007), No. 2, pp. 307–316

The Influence of Niobium Microalloying on Austenite Grain Coarsening Behavior of Ti-modified SAE 8620 Steel K. A. ALOGAB,1) D. K. MATLOCK,2) J. G. SPEER2) and H. J. KLEEBE3) 1) Department of Mechanical Technology, Buraydah College of Technology, Buraydah, Saudi Arabia. E-mail: [email protected] 2) Advanced Steel Processing and Products Research Center, Department of Metallurgical and Materials Engineering, Colorado School of Mines, Golden, CO 80401. E-mail: [email protected], [email protected] 3) Institute for Applied Geosciences, Schnittspahnstrasse 9, Technical University of Darmstadt, D-64287 Darmstadt, Germany. E-mail: [email protected] (Received on July 20, 2006; accepted on November 7, 2006 )

The potential for suppressing unacceptable austenite grain growth during carburizing by Nb microalloying additions in the range of 0.02 to 0.11 wt% to a Ti-modified SAE 8620 carburizing steel were evaluated. Alloys, were designed based on fundamental equilibrium thermodynamic analyses, as part of an extensive study on the effects of alloy composition, thermomechanical history, and pseudo-carburizing conditions on austenite grain coarsening behavior. Laboratory samples were produced to simulate both conventional hot rolling and controlled rolling practices designed to produce different initial precipitate distributions. Pseudocarburizing heat treatments, i.e. without a carburizing atmosphere, were performed in the temperature range of 950 to 1 100°C for holding times of 30 to 360 min. Precipitate distributions, including size, number density, morphology, distribution, and chemical composition in selected samples from the as-rolled and pseudo-carburized conditions were evaluated with transmission electron microscopy on extraction replicas. Results showed that increasing Nb additions to the Ti-modified SAE 8620 steel restrained austenite grain coarsening, and increased the grain coarsening time, especially at temperatures below 1 050°C. The Nb-free (Ti-modified) steel yielded either severely duplex grain structures or pseudo-normal grain growth (with very large mean grain diameter). However, holding a Ti–Nb-modified steel (e.g. 0.06 Nb wt%) at 950°C for 6 h or at 1 000°C for 4 h. produced fine and uniform austenite grain structures (with a mean grain diameter less than 20 m m). The finer grain sizes observed in the Ti–Nb-modified steels were due to the presence of Nbrich precipitates that hinder austenite grain coarsening, and precipitate distributions and grain growth behaviors are also influenced by the steel rolling history. The results indicate that Nb can successfully be used to suppress grain growth in carburizing steels. KEY WORDS: niobium microalloying; carburizing; precipitation; abnormal grain growth.

1.

tional time and expensive processing steps. Thus, several recent studies have considered alloy and process modifications designed to produce suitable precipitate distributions to suppress austenite grain coarsening at carburizing temperatures.7–17) Previous studies have considered processing to control aluminum nitride precipitates8) and alloying to control precipitates based on titanium and nitrogen additions,7) niobium,9) or titanium plus niobium additions.10–16) While the previous studies have illustrated the potential of utilizing precipitates to limit austenite grain coarsening during carburizing, there have been no systematic studies designed to provide the necessary fundamental information to optimize precipitate distributions to minimize grain coarsening during carburizing. The present work investigates the effects of niobium additions and thermomechanical history on the austenite grain coarsening behavior in a Ti-modified SAE 8620 steel, designed for use in forged and carburized gears. Niobium was considered because of its potential to form a more controllable grain refiner (NbC), in combination with potential benefits of TiN dispersions as developed by Davidson et al.7)

Introduction

There is substantial interest in the development and use of a new generation of grain-refined carburizing steels for automotive applications with a goal of increasing carburizing temperatures. In conventional gas carburizing processing, carburizing is typically conducted at temperatures between 870 and 980°C for times from a few hours to a day depending on the desired case properties.1) Increasing carburizing temperature, however, has long been known to significantly shorten carburizing cycle times.2) One important obstacle that limits the use of high carburizing temperatures is excessive austenite grain coarsening that may occur during carburizing processes. Carburizing of conventional steel grades above 950°C typically requires post-carburizing heat treatments to produce a final microstructure with a refined prior austenite grain size.3) Fine carburized case microstructures are desirable as they have been shown to result in improved mechanical properties, particularly fatigue resistance.4–6) Post-carburizing heat treatments to refine the prior austenite grain size are undesirable as they require addi307

© 2007 ISIJ

ISIJ International, Vol. 47 (2007), No. 2

burizing temperatures. Niobium additions in the range of 0.02 to 0.3 (wt%) were considered. Calculations, based on the solubility data of Narita,19) addressed the weight percent of NbC precipitates that formed prior to rolling and prior to carburizing, the amount of dissolved niobium present while reheating to the carburizing temperature, and the amount of the additional NbC precipitates that might form during carburizing. As expected, the volume fraction of NbC precipitates was predicted to increase with increasing niobium and carbon levels. However, the volume fraction of the additional NbC precipitates, which may form during carburizing, has a limit. To clarify this point, consider two alloys containing 0.1 and 0.3Nb with 0.2C (wt%) carburized at 1 100°C, as shown in Fig. 2. From the solubility plot displayed in Fig. 2, evaluation of these two alloys at the specified temperature predicted significant differences in the initial precipitate volume fractions, as indicated by “D fppt prior to carburizing.” With an increase in carbon content during carburizing, indicated here for a carbon content of 0.6 wt% chosen to approximate the average carbon content in a typical carburized layer, the NbC fraction would increase only slightly and the difference between the two alloys (indicated “D fppt during carburizing”) would be negligible. The probability of forming additional NbC precipitates during the carburizing process depends primarily on the amount of Nb remaining in solution prior to the carbon enrichment step in the process. Figure 3 shows the computed equilibrium amounts of TiN and NbC at various austenitizing temperatures (in the absence of an increased carbon potential associated with the carburizing process) in a 0.03Ti base alloy and three Nb-modified alloys with niobium concentrations between 0.02 and 0.13Nb (wt%). Consistent with Fig. 1, this figure shows that TiN precipitates are stable even at temperatures approaching 1 300°C, whereas the amount of Nb as NbC precipitates would decrease with increasing temperature. Based on the equilibrium thermodynamic calculations, three experimental steels with Nb additions (low, medium and high) were designed based on the Ti-modified SAE 8620 steel (with a Ti : N ratio of approximately 3.42, i.e. a stoichiometric ratio). These steels included (in wt%) 0.02, 0.06, and 0.13Nb with 0.03Ti and 80 ppm N. It was anticipated that the experimental alloys would produce marked differences in austenite grain coarsening behavior when heat treated or carburized. The maximum Nb level was lim-

2. Alloy and Process Design 2.1. Thermodynamic Basis for Alloy Selection Experimental alloys, intended to assess the ability of niobium additions to suppress austenite grain coarsening at elevated temperatures, were developed based on fundamental equilibrium thermodynamic analyses of niobium-carbide precipitate formation. Following the previous work of Davidson et al.,7) which showed that a maximum retardation effect on austenite grain coarsening at carburizing temperatures was observed in a Ti-modified SAE 8620 steel with a Ti : N ratio close to stoichiometric (3.42 by weight), the experimental alloys were developed based on the following assumptions: i) Titanium added stoichiometrically to the SAE 8620 steel will precipitate N as TiN; ii) In the Ti-modified SAE 8620 steel, added niobium will precipitate as NbC; and iii) Maximum temperatures experienced during solid-state processing are below about 1 300°C. With these assumptions, the desired niobium levels were determined based on solubility product analyses for pure NbC. For a Ti-microalloyed SAE 8620 steel containing, in wt%, 0.2C, 0.008N, and 0.03Ti (referred to as the base alloy) and different niobium concentrations, the amount of niobium-carbide precipitates present before and after carburizing treatments were predicted based on equilibrium thermodynamic calculations. To provide background information for assessment of appropriate niobium additions for considerations in this study, Fig. 1, developed based on solubility constants of Matsuda and Okumura,18) presents a titanium-nitride solubility diagram for three different temperatures (900, 1 200, and 1 300°C) along with the composition of the base alloy and a dashed line indicating stoichiometric combinations of titanium and nitrogen. At all carburizing temperatures of interest, Fig. 1 indicates that most of the nitrogen in the base alloy precipitates as TiN and, thus, the first assumption above is reasonable. To assess the effects of precipitate formation in alloys with controlled niobium contents, calculations were made assuming that niobium additions lead to NbC, rather than a complex TiNb(C, N) carbonitride. The results of this analysis are summarized below and were used to design specific alloys for an experimental study on austenite grain growth at car-

Fig. 1. TiN solubility isotherms at various temperatures, demonstrating the levels of solute Ti and N in equilibrium with TiN. In the base alloy, titanium and nitrogen are largely combined in the form of TiN. Calculations are based on equilibrium solubility constants reported by Matsuda and Okumura.18)

© 2007 ISIJ

Fig. 2. NbC solubility diagram at 1 100°C, illustrating differences in the amount of NbC before and after carburizing, for two different Nb levels. Calculations are based on equilibrium solubility constants reported by Narita.19)

308

Fig. 3. Calculated equilibrium weight percents of NbC (solid lines) and TiN (dashed line) in Ti–Nb modified SAE 8620 steels, with different niobium levels.

ISIJ International, Vol. 47 (2007), No. 2

Fig. 4. (a) Computed equilibrium amount of dissolved Nb versus temperature in a 0.2C steel containing different Nb levels, (b) calculated amount of NbC precipitates that may form during carburizing (with carbon concentration of 0.6 wt%).

Fig. 5. NbC solubility isotherms in austenite at two temperatures illustrating the difference between the equilibrium amounts of NbC at the two finish-rolling temperatures (solubility data of Narita19)).

Table 1. Chemical compositions of experimental steels (in wt%, N and O in ppm).

ited to approximately 0.13 wt%, as solubility considerations indicated that higher levels of Nb could not be taken into solution at any temperatures applicable to bar processing in the solid-state. Further details on the alloy design are provided elsewhere.11) For steels including the range of Nb levels of the experimental steels, Fig. 4 shows the levels of solute Nb present at the carburizing temperature. Figure 4(a) shows the computed amount of dissolved Nb at different reheat temperatures for a 0.2C level (representing the core, or the case prior to carbon enrichment). Based on this figure the maximum amount of Nb that can be dissolved in a 0.2C (wt%) steel at temperatures up to 1 200°C is approximately 0.06 (wt%). Figure 4(b), shows the amount of ‘additional’ NbC that may form in the case region, assuming a near-surface carbon concentration as 0.6 wt. Based on Fig. 4(b), differences in the amount of NbC that may form during carburizing in steels containing total Nb additions in the range of 0.02 to 0.13 (wt%) are negligible at carburizing temperatures below 1 100°C, and become more substantial at higher temperatures. While the influence of the carbon potential on austenite grain-growth behavior is considered in the pseudo-carburizing results presented here, the calculations provide insights into the precipitation behavior expected in the experimental steels in future carburizing studies.

1 300°C, respectively and these temperatures were used to guide the selection of a soaking temperature for the experimental study discussed below. A consideration of the effects of finishing temperature, for both HR and CR processes, was used to develop the final processing schedule. The influence of the finishing temperature on the NbC precipitation may be understood for the 0.06Nb alloy by considering Fig. 5. This figure shows the solubility curves for NbC at 850 and 1 100°C, together with the equilibrium amount of NbC at the two finish-rolling temperatures. Based on Fig. 5, the equilibrium amount of NbC precipitate in the CR Nb-containing alloy was predicted to be significantly higher than that in the HR material at the respective finish-rolling temperatures. Similar predictions apply to the other two Nb alloys. In addition, the lower processing temperatures associated with typical CR processing histories should result in a distribution of finer precipitates in the austenite. Similarly, it is reasonable to expect that differences in NbC precipitation in ferrite might also occur, since the amount of Nb prior to transformation is so different for the two finish rolling temperatures. Therefore, the two thermomechanical schedules were applied to assess the importance of processing on microstructural evolution in the hot-rolled condition and after pseudo-carburizing.

2.2. Thermomechanical Processing Conditions To assess the influence of starting precipitate distributions in the ‘as-rolled’ Nb-containing alloys on austenite grain coarsening, two rolling processes, designed to simulate conventional hot rolling (HR) and controlled rolling (CR) processes, were used. the calculated NbC dissolution temperatures for the 0.02Nb, 0.06Nb, and 0.13Nb experimental alloys were approximately 1 100, 1 200, and

3.

Experimental Materials and Procedure

Three experimental Ti–Nb alloys were prepared as 45 kg lab cast heats with the chemical compositions given in Table 1 and the alloys are referred to by their nominal niobium contents. The actual Nb content of 0.109 wt% for the high Nb alloy was slightly less than the 0.13 wt% specified in the design section above. A Nb-free steel was available 309

© 2007 ISIJ

ISIJ International, Vol. 47 (2007), No. 2

from a previous study7) and was included in this study for comparison. The Ti–Nb alloy ingots were reheated to 1 230°C, forged, and then air-cooled to room temperature. The billets were machined to a thickness of 63.5 mm, soaked at 1 260°C (a temperature predicted to produce complete precipitate dissolution in the 0.109 wt% Nb alloy) for about one hour, and subjected to two different rolling schedules, simulating conventional hot-rolling (HR) or controlled-rolling (CR). For the HR schedule, six single-pass reductions of 22, 13, 15, 18, 22, and 29% were used. After the third pass, the plates were reheated for 10 min at 1 230°C and then rolled in three passes to a final thickness of 19 mm. The last rolling pass in the HR process was applied at approximately 1 100°C. For the CR schedule, nine single-pass reductions of 22, 13, 15, 9, 9, 11, 12, 14 and 15% were used. The first four reductions occurred above 1 038°C. After the fourth pass, plates in the CR schedule were air-cooled to approximately 927°C and subsequently rolled in five passes to a final thickness of 19 mm, with a finish rolling temperature of approximately 850°C. The processing history for the Nb-free steel was presented elsewhere.7) To evaluate the effectiveness of the alloy modifications and thermomechanical histories on the austenite grain coarsening behavior and precipitate distributions, 2015 10 mm3 specimens were pseudo-carburized (i.e. austenitized in a non-carburizing atmosphere) at temperatures between 950 and 1 100°C for times between 30 and 360 min and ice-water quenched. Specimens were heated to the pseudo-carburizing temperatures at a rate of approximately 145°C min1 (as measured with an embedded thermocouple), and indicated times represent the holding time after reaching the specified temperature. Austenite grain boundaries in the pseudo-carburized samples were revealed on standard metallographic mounts etched in a saturated aqueous picric acid solution containing a wetting agent (teepol) and HCl. Details on the etching procedures are provided elsewhere.11) Austenite grain sizes were measured on printed micrographs by the line intercept method according to ASTM E112. Grain boundary intercepts with a 500 mm line were counted in ten fields per specimen. Carbon extraction replicas were used to examine precipitate characteristics, including size, number density, morphology, chemical composition, and distribution using a scanning transmission electron microscope (TEM) operating at an accelerating voltage of 120 kV. Analyses of precipitate distributions were performed on specimens selected from the as-received and pseudo-carburized conditions. For this TEM study, pseudo-carburized samples were selected based on austenite grain coarsening behavior observed in light optical microscopy. For each sample evaluated in the TEM, at least three replicas were examined, and for each replica at least seven different fields of view were imaged. At least 200 precipitates, excluding precipitates larger than 100 nm, were measured to characterize the precipitate size distribution. Precipitate number densities were estimated by counting observed particles in a nominal total area of 36.7 m m2 (summed from multiple fields of view), and results are expressed as number of particles per square micron. The prior-austenite grain structures were characterized with specific terms that highlighted critical characteristics. These characteristics include: (a) the condition that corresponded to the Initial observation of Abnormal Grain Growth (IAGG), i.e. the appearance of one large or a few large grains in a relatively equiaxed fine grained matrix; (b) © 2007 ISIJ

Fig. 6. Electron micrographs from extraction replicas showing the effects of thermomechanical history on precipitation behavior in the as-rolled 0.06Nb steel: (a) HR and (b) CR.

the Grain Coarsening Time, tGC, the time after which IAGG was observed at a given temperature; (c) the condition of Abnormal Grain Growth (AGG), i.e. the presence of a duplex grain structure; and (d) Normal Grain Growth (NGG) or “pseudo-normal” grain growth, i.e. grain growth without a distinct duplex grain structure. 4. Results 4.1. Precipitate Distributions in the As-rolled Steels Dispersed cuboidal precipitates with sizes up to 5 m m were observed by light microscopy in the as-rolled materials.11) While not shown here, the large cubic particles were interpreted as TiN formed either in the liquid state or during solidification.20–22) It was anticipated that these precipitates were too large to significantly limit austenite grain growth, and thus were excluded from the precipitate distribution analyses presented below. Transmission electron micrographs from the extraction replicas of the as-rolled 0.06Nb steel, for both the HR and CR conditions are shown in Fig. 6. A higher number density of fine precipitates is observed in the CR steel (Fig. 6(b)), and similar effects of rolling were anticipated for the other two Nb-containing steels (0.02Nb and 0.1Nb). Figure 7 shows the precipitate size distributions obtained for each sample used to obtain the electron micrographs presented in Fig. 6. Figures 6 and 7 confirm that the CR steel possessed both a finer average precipitate size and a higher number density of fine precipitates. 4.2. Austenite Grain Coarsening Behavior Figure 8 presents light optical micrographs of the three HR steels austenitized for 90 min at 950°C (Figs. 8(a) to 8(c)), 1 000°C (Figs. 8(d) to 8(f)), and 1 050°C (Figs. 8(g) to 8(i)) arranged to facilitate direct comparisons of the effects of Nb content on austenite grain coarsening behavior after austenitizing at different temperatures. For a specific austenitizing temperature, each column shows the effects of increasing Nb addition. At 950°C and 1 000°C the micrographs show that an increase in Nb content results in a transition in austenite grain structure from a duplex structure characteristic of AGG to a fine and uniform grain structure characteristic of NGG. At 1 050°C abnormal grain growth was observed in all samples although AGG was not observed in the higher Nb steels for annealing times less than for the samples shown in Fig. 8.11) Measurements of the mean austenite grain size were used to quantify the effects of time and temperature on the effects of pseudo-carburizing heat treatments on austenite grain coarsening behavior, and the results are summarized in Fig. 9 for the HR steels and in Fig. 10 for the CR steels. 310

ISIJ International, Vol. 47 (2007), No. 2

Fig. 7. Precipitate size distribution for the asrolled 0.06Nb steel: (a) HR and (b) CR.

Fig. 8. Typical examples of light optical micrographs showing austenite grain structures for selected HR steels after pseudo-carburizing for 90 min at different temperatures: (a) to (c) 950°C; (d) to (f) 1 000°C; and (g) to (i) 1 050°C.

The base steel (Nb-free) was hot rolled7) and thus its grain coarsening behavior is included with the HR steels in Fig. 9. In each figure, the mean austenite grain diameter is plotted versus pseudo-carburizing time for a specific alloy/temperature combination. Error bars represent 95% confidence intervals about the mean grain diameters. The data presented in Figs. 9 and 10 show three distinct characteristics depending on alloy composition and pseudocarburizing temperature: (a) fine and uniform grain sizes characteristic of NGG, with a mean grain diameter less than approximately 20 m m which depends slightly on austenitizing time; (b) regions of significant grain coarsening with time associated with AGG (i.e. observations of duplex grain structures); and (c) regions of uniform or “pseudo-uniform” grain size, but with very large mean grain diameters which depend slightly on austenitizing time, and which have a characteristics of NGG or “pseudonormal” grain growth. Figures 9 and 10 reveal several important observations about the effects of Nb content and processing history on austenite grain coarsening behavior in microalloyed-modified carburizing steels. In comparison to the base steel (Nbfree), Nb additions suppress austenite grain coarsening and the effect increases with Nb content (e.g. Fig. 9(a)). For both the HR and CR steels pseudo-carburized at temperatures between 950 and 1 050°C, the susceptibility to AGG decreased (i.e. tGC, increased) with an increase in Nb content or a decrease in pseudo-carburizing temperature. For all conditions, temperature primarily altered the conditions for the onset of AGG as opposed to the growth rate during

NGG. Figure 9(d) shows that significant AGG occurred in the HR steels pseudo-carburized at the highest temperature, 1 100°C, such that the mean grain diameters in the Nb-containing steels were greater than in the base steel. For all heat treatment conditions, the CR Nb-modified steels exhibited finer mean austenite grain diameters than the corresponding HR Nb-modified steels. For example, for steels pseudo-carburized at 1 050°C, the maximum mean grain diameter for the CR steel was about 60 m m (Fig. 10(c)) while for the HR steel it was about 120 m m (Fig. 9(c)). The effects of alloy composition and pseudo-carburizing conditions (time and temperature) on austenite grain coarsening are interpreted based on precipitate distributions as described below. 4.3. Precipitate Distributions after Pseudo-carburizing Six pseudo-carburized samples, selected based on specific differences in grain growth response shown in Figs. 9 and 10, were examined in the TEM. Carbon extraction replicas were prepared and used to investigate precipitate characteristics, including precipitate size, precipitate number density, morphology, and distribution for the selected conditions. Six samples, selected for TEM analysis included the following comparisons: (a) HR samples of the three Nb-modified steels pseudocarburized at 1 050°C for 60 min and (b) CR samples of the 0.06Nb steel pseudo-carburized at 950°C for 60 min and at 1 050°C for 30 and 60 min. Figure 11 shows selected TEM micrographs that illustrate the precipitate morphologies and size distributions. 311

© 2007 ISIJ

ISIJ International, Vol. 47 (2007), No. 2

Fig. 9. Mean austenite grain diameter versus pseudo-carburizing time for the HR Nb-containing steels along with the base steel (Nb-free) at different temperatures: (a) 950°C; (b) 1 000°C; (c) 1 050°C; and (d) 1 100°C.

Fig. 10. Mean austenite grain diameter versus pseudo-carburizing time for the CR Nb-containing steels at different temperatures: (a) 950°C; (b) 1 000°C; (c) 1 050°C; (d) 1 100°C.

© 2007 ISIJ

312

ISIJ International, Vol. 47 (2007), No. 2 Table 2. Particle number density and average size after different conditions.

Fig. 11. Electron micrographs from extraction replicas showing precipitate distributions in hot rolled (HR) and controlled rolled (CR) steels: (a) HR 0.02Nb, 60 min at 1 050°C; (b) HR 0.06Nb, 60 min at 1 050°C; (c) HR 0.1Nb, 60 min at 1 050°C; (d) CR 0.06Nb, 60 min at 950°C; (a) CR 0.06Nb, 30 min at 950°C; and (d) CR 0.06Nb, 60 min at 1 050°C.

Fig. 12. Correlation of precipitate size to the calculated peak intensity ratios (expressed as percent) based on Nb and NbTi Ka peaks: (a) CR 0.06Nb steel after different conditions; and (b) different steels pseudo-carburized at 1 050°C for 60 min.

particle number density in the as-rolled condition is less than 1/2 of that observed in the reaustenitized samples (see Table 2). Increasing austenitizing time at 1 050°C from 30 to 60 min yielded a slight reduction in the Nb/(TiNb) ratio for precipitate sizes less than 40 nm, as can be seen in Fig. 12(a). This reduction in Nb/(TiNb) ratio is attributed to the coarsening of Nb-rich precipitates, a process itself which requires dissolution of Nb particles. Figure 12(b) shows the composition dependence on particle size in two HR steels (0.02Nb and 0.1Nb) austenitized at 1 050°C for 60 min. Data for the CR 0.06Nb steel austenitized at 1 050°C for 60 min, which are included as part of Fig. 12(a), are reproduced in Fig. 12(b) for comparison. Based on Fig. 12(b), the ratio of Nb/(TiNb) in the examined particles appears to be dependent on the original Nb/(TiNb) ratio in the examined steels. Furthermore, the increase in particle size above about 40 nm sharply reduced the Nb/(TiNb) ratio, as can be seen in Figs. 12(a) and 12(b). This latter conclusion confirms that austenite grain coarsening in the examined steels is predominantly controlled by the presence of Nb-rich precipitates (due to their small sizes) and further suggests that TiN-rich particles (due to their initial large sizes) have a minor effect on controlling grain coarsening.

From multiple images similar to those shown in Fig. 11, quantitative measurements of the number densities and precipitate size distributions were obtained.11) These results are summarized in Table 2 and show that, for samples pseudocarburized at 1 050°C for 60 min, an increase in Nb content increases both the particle number density and average particle size. A comparison of the HR and CR 0.06Nb samples pseudo-carburized at 1 050°C for 60 min. shows that the CR condition exhibits a significantly higher particle number density and finer average particle size. Precipitate chemical compositions were analyzed using energy dispersive spectroscopy (EDS). Sample conditions that were analyzed included: CR 0.06Nb in the as-received condition; CR 0.06Nb pseudo-carburized at 1 050°C for 30 and 60 min; and HR 0.02Nb and 0.1Nb pseudo-carburized at 1 050°C for 60 min. Precipitate chemical compositions, which are detailed elsewhere,11) were estimated based on the ratio of Nb to NbTi Ka peak intensities and plotted versus particle size in Fig. 12. Figure 12(a) summarizes the composition dependence on particle size for the CR 0.06Nb steel in the as-rolled and heat treated conditions. From this figure, it can be seen that reaustenitizing treatments increase the Nb/(TiNb) ratio significantly in comparison to the as-rolled condition. The Nb/(TiNb) ratio decreases with increasing austenitizing time and/or particle size. The higher Nb/(TiNb) ratio values in the reaustenitized samples, in comparison to the as-rolled condition, suggest additional Nb-rich precipitates formed during the heat treatment process. These observations are consistent with the precipitate number density quantifications, which showed that the

5.

Discussion

The results presented above illustrate the potential for Nb microalloying to suppress austenite grain coarsening in steels subjected to higher temperatures applicable to plasma and vacuum carburizing operations. The experimental observations are consistent with previous research on the effects of second phase particles on normal and abnormal grain growth.17,23–29) The combined effects of alloy content 313

© 2007 ISIJ

ISIJ International, Vol. 47 (2007), No. 2

and thermal processing on the resulting austenite grain structure are complex and reflect competition between the driving forces for normal/abnormal grain growth and pinning forces of precipitates as controlled by precipitate type and size distribution.17,30–43) Abnormal grain growth is an important consideration in steels with pinning particles, since AGG can lead to austenite grains that are significantly greater in size than those observed in plain carbon steels where NGG is manifested at all temperatures.32,36,43,44) In the steels investigated here, alloy content, casting process, rolling history, and pseudo-carburizing conditions all contribute to the development of the precipitate distribution involved in suppressing austenite grain growth. Since the TiN distribution, assumed to be similar in each steel, developed primarily in the initial casting and hot rolling procedures, differences in the grain growth response are attributed to the presence of NbC precipitates and the effects of pseudo-carburizing time and temperature on NbC precipitate distributions and grain boundary mobility. Austenite grain growth in the Nb-free base steel exhibited either AGG (duplex microstructure) or pseudo-NGG (with a very large mean grain diameter). Abnormal grain growth in the Nb-free base steel was only observed in samples pseudo-carburized for short times at 950°C whereas the pseudo-NGG mode was observed at longer times at the lowest pseudo-carburizing temperature and for all the examined times at the higher pseudo-carburizing temperatures; true NGG (i.e. the presence of fine and uniform austenite grains) was not observed for any annealing condition. The absence of fine and uniform austenite grain structures after heat treating at the relatively lower temperatures confirms that the Nb-free base steel exhibited an insufficient number of effective TiN particles (i.e. particles below a critical size required to restrain grain growth) in the as-

rolled condition. This conclusion is consistent with the previous work of Davidson,44) who found that the base steel (Nb-free steel) exhibited large TiN precipitates (approximately 50 nm), in the as-rolled condition. The combined effects of processing history and Nb levels were compared by assessing average austenite grain sizes and time to initiate abnormal grain growth (tGC) at each pseudo-carburizing temperature. Table 3 summarizes tGC values obtained from the grain growth data shown in Figs. 9 and 10. Addition of Nb to the base steel increased the grain coarsening time, and for all the Nb-containing steels, the grain coarsening time decreased with an increase in temperature. For the 0.02Nb and 0.06Nb steels pseudocarburized at temperatures less than 1 050°C, the grain coarsening times for the HR steels were slightly greater than for the CR steels. Figure 13 summarizes the effects of Nb additions on the mean austenite grain size for samples pseudo-carburized at 950, 1 000, 1 050, and 1 100°C for different times. This figure shows that an increase in Nb content resulted in different grain growth responses, and the CR steels produced finer mean austenite grain sizes for all the examined pseudo-carburizing conditions as compared to the HR steels. Differences in rolling history are most evident in Table 3. Summary for the grain coarsening time (tGC), in min.

Fig. 13. Mean austenite grain diameter as a function of niobium addition at different temperatures: (a) 950°C; (b) 1 000°C; (c) 1 050°C; and (d) 1 100°C. Annealing times and rolling schedules are given in the plots.

© 2007 ISIJ

314

ISIJ International, Vol. 47 (2007), No. 2

samples pseudo-carburized at the higher temperatures (e.g. at 1 100°C), where the austenite grain size in the CR samples is less than 1/2 that of the HR samples. Based on the current results, the effects of rolling history on grain coarsening behavior can be stated as follows: (i) grain coarsening time, tGC, is shorter in the CR steel than in the corresponding HR steel; and (ii) after tGC, (IAGG), the duplex grain structure in the HR steel grows much faster than in the corresponding CR steel. Figures 6 and 7 show that controlled rolling produced a higher density of fine precipitates, a consequence in part of the effects of the lower finish rolling temperature on the solubility of Nb (e.g. Fig. 5) and reduced recovery rates which potentially lead to higher dislocation densities that provided a high number of heterogeneous nucleation sites in the austenite.45–48) The observed transitions in grain coarsening behavior with Nb content and temperature correlate directly to the effects of alloy content and temperature on the stability of the restraining particles. With an increase in Nb content, the precipitate volume fraction and the critical precipitate size required to pin austenite grain boundaries increase.49) On the other hand, increasing the temperature decreases the stability of a precipitate, which leads to higher coarsening/dissolution rates, consistent with Ostwald ripening theory.50) Pseudo-carburizing the HR and CR 0.02Nb steels at 950°C for times less than their corresponding grain coarsening times and the HR and CR 0.06Nb and 0.1Nb steels at all investigated times at 950 and 1 000°C produced uniform and fine austenite grain structures. Suppressing austenite grain coarsening in these steels at times below the tGC can be attributed primarily to the presence of Nb-rich precipitates that effectively hinder abnormal grain growth until the driving force for grain growth exceeds the pinning force exerted by the Nb-rich precipitates. Figure 11(d) shows that the majority of precipitates in the CR 0.06Nb specimen pseudo-carburized at 950°C for 60 min were less than 20 nm, and thus were able to stabilize the fine and uniform austenite grain structure observed for this heat treatment condition. At higher pseudo-carburizing temperatures, the coarsening rate of Nb-rich precipitates is accelerated, which allows the release of selected grains in a relatively short time, resulting in a significant reduction in the tGC. Abnormal grain growth occurs when the combination of particle size, r, and volume fraction, f, is insufficient to pin austenite grain boundaries (e.g. low f/r ratio), allowing selected grains to grow unstably.35) TEM observations showed the existence of fine Nb-rich precipitates in the HR 0.02Nb steel that was pseudo-carburized at 1 050°C for 60 min (Fig. 11(a)) but the volume fraction of the available effective precipitates was apparently insufficient to retard austenite grain boundaries and abnormal grain growth occurred in a relatively short time at 1 050°C. The onset of abnormal grain growth was accelerated by decreasing the precipitate volume fraction (due to dissolution) and/or increasing pinning particle sizes (due to coarsening), resulting in unpinning of some grain boundary regions. When the restraining particles exceeded a critical size, growth of certain grains was favored49) promoting the development of duplex grain structures do to abnormal grain growth as illustrated by the 0.02Nb steel annealed above 950°C, and by the 0.06Nb and 0.1Nb steels annealed above 1 000°C. For example for the HR 0.06Nb steel, an increase in the pseudo-carburizing temperature from 1 000 to 1 050°C reduced the tGC from above 240 to 30 min. Furthermore, heat treating the 0.06Nb and 0.1Nb steels (both rolling schedules) at 1 100°C resulted in the development of uniform large grains, an obser-

Fig. 14. Microstructure map based on austenite grain growth modes for the CR 0.02Nb steel. Solid and opened circles represent the experimental data.

vation attributed to the rapid coarsening and/or dissolution of the NbC precipitates. The absence of a duplex grain structure (AGG), such as that shown in Fig. 8(i), for a specific heat treatment condition does not necessarily preclude the earlier occurrence of AGG. It simply indicates that, at the treatment conditions where AGG does not exist, the coarse grains consumed the fine grains at shorter times leading to a structure that appeared as being due to normal grain growth. Based on the results of this study, interactions between pseudo-carburizing time and temperature can be summarized in microstructure maps for each Nb addition and process history. As an example, Fig. 14 presents a map for the CR 0.02Nb steel. This figure clearly shows three distinct grain growth modes in the CR 0.02Nb steel depending on heat treatment time and temperature. In the regime at short times and/or low temperatures, the extent of grain growth is very small (NGG), which is attributed to the presence of a sufficient number of effective Nb-rich precipitates. With an increase in time or temperature the effective precipitate dispersion becomes unstable due to Ostwald ripening, the condition for IAGG is approached and abnormal grain growth develops. At higher heat treatment times or temperatures, the excessive coarsening and/or dissolution of the effective Nb-rich precipitates takes place and leaves behind only the more stable and coarse TiN precipitates resulting in a more uniform large grain size structure characteristic of pseudo-NGG with very large grain diameters. The study presented here concentrated on pseudo-carburizing heat treatments. However, thermodynamic predictions indicate that in carburizing atmospheres, the additional precipitation due to the higher carbon contents associated with the carburized layer is limited at temperatures up to about 1 100°C. The higher matrix carbon content may further enhance NbC coarsening resistance, however, since the matrix would be expected to be even further depleted in solute-Nb, which is an important factor influencing precipitate coarsening kinetics. While additional research is required to confirm the effects of carbon content on the precipitation behavior, this study clearly illustrates the potential for producing new carburizing steels with an increased resistance to unacceptable grain growth during carburizing. Additionally, the influence of spheroidizing treatments and other pre-carburizing process variants will require consideration in future studies. 315

© 2007 ISIJ

ISIJ International, Vol. 47 (2007), No. 2

6.

Res., 8 (2005), No. 4, 453. 13) K. A. Alogab, D. K. Matlock and J. G. Speer: Proc. of the Fourth Saudi Tech. Conf. & Exhib. (STCEX2006), vol. 3, Gotevot, Riyadh, Saidi Arabia, (2006), 183. 14) S. Trute, W. Bleck, and C. Klinkenberg: Proc. of Int. Conf. on New Developments in Long and Forged Products: Metallurgy and Applications, ed. by J. G. Speer, E. B. Damm and C. V. Darragh, AIST, Warrendale, PA, USA, (2006), 161. 15) T. Murakami, H. Hatano and H. Yaguchi: Proc. of Int. Conf. on New Developments in Long and Forged Products: Metallurgy and Applications, ed. by J. G. Speer, E. B. Damm and C. V. Darragh, AIST, Warrendale, PA, USA, (2006), 99. 16) C. Klinkenberg and S. G. Jansto: Proc. of Int. Conf. on New Developments in Long and Forged Products: Metallurgy and Applications, ed. by J. G. Speer, E. B. Damm and C. V. Darragh, AIST, Warrendale, PA, USA, (2006), 135. 17) T. Gladman: Grain Size Control, Institute of Materials, Minerals & Mining, London, England, (2004), 99. 18) S. Matsuda and N. Okumura: Trans. Iron Steel Inst. Jpn., 18 (1978), 198. 19) K. Narita: Trans. Iron Steel Inst. Jpn., 15 (1975), 145. 20) S. F. Medina, M. Chapa, P. Valles, A. Quispe and M.I. Vega: ISIJ Int., 39 (1999), 930. 21) D. C. Houghton, G. C. Weatherly and J. D. Embury: Processing of Microalloyed Austenite, ed. by A. J. DeArdo, G. A. Ratz and P. J. Wray, The Metall. Soc. of AIME, (1981), 267. 22) S. Akhlaghi and D. G. Ivey: Can. Metall. Q., 41 (2002), 111. 23) O. O. Miller: Trans. Am. Soc. Met., 43 (1951), 260. 24) L. J. Cuddy and J. C. Raley: Metall. Trans. A, 14A (1983), 1989. 25) H. Adrian and F. B. Pickering: Mater. Sci. Technol., 7 (1991), 176. 26) MD. M. A. Bepari: Metall. Trans. A, 20A (1989), 13. 27) J. M. Cabrera, A. Al Omar and J. M. Prado: J. Mater. Sci., 31 (1996), 1303. 28) G. Cai and J. D. Boyd: Grain Growth in Polycrystalline Materials III, ed. by H. Weiland, B. L. Adams and A. D. Rollett, TMS, Warrendale, PA, (1998), 579. 29) K. He and T. N. Kaker: Mater. Sci. Eng. A., A256 (1998), 111. 30) C. S. Smith: Trans. AIME, 175 (1948), 15. 31) T. Gladman and F. B. Pickering: J. Iron Steel Inst., 205 (1967), 653. 32) T. Gladman and D. Dulieu: Met. Sci., 8 (1974), 167. 33) T. Gladman: Recrystallization and Grain Growth of Multi-Phase and Particle Containing Materials, ed. by N. Hansen, A. R. Jones and T. Leffers, RISØ, Roskilde, Denmark, (1980), 183. 34) M. F. Ashby: Recrystallization and Grain Growth of Multi-Phase and Particle Containing Materials, ed. by N. Hansen, A. R. Jones and T. Leffers, RISØ, Roskilde, Denmark, (1980), 325. 35) T. Gladman: Grain Growth in Polycrystalline Materials I, ed. by G. Abbruzzese and P. Brozzo, Trans. Tech. Pub., Zurich, Switzerland, (1992), 113. 36) R. Elst, J. V. Humbeeck and L. Delaey: Acta Metall., 36 (1988), 1723. 37) I. Andersen and Ø. Grong: Acta Metall. Mater., 43 (1995), 2673. 38) I. Andersen, Ø. Grong and N. Ryum: Acta Metall. Mater., 43 (1995), 2689. 39) P. R. Rios: Acta Metall., 35 (1987), 2805. 40) P. A. Manohar, D. P. Dunne, T. Chandra and C. R. Killmore: ISIJ Int., 36 (1996), 194. 41) T. Gladman, G. Fourlaris and M. Talafi-Noghani: Mater Sci. Technol., 15 (1999), 1414. 42) T. Gladman: The Physical Metallurgy of Microalloyed Steels, Inst. of Materials, London, England, (1997), 176. 43) M. L. Santella and A. J. DeArdo: Thermomechanical Processing of Microalloyed Austenite, ed. by A. J. DeArdo, G. A. Ratz and P. J. Wray, The Metall. Soc. AIME, Warendale, PA, (1981), 88. 44) S. G. Davidson: M. S. Thesis, Colorado School of Mines, Golden, CO, (2001). 45) J. G. Speer and S. S. Hansen: Metall. Trans. A, 20A (1989), 25. 46) A. Le Bon, J. Rofes-Vernis and C. Rossard: Met. Sci., 9 (1975), 36. 47) J. J. Jonas and I. Weiss: Met. Sci., 13 (1979), 238. 48) Dutta and C. M. Sellar: Mater. Sci. Technol., 3 (1987), 197. 49) T. Gladman: Proc. R. Soc. (London), A294 (1966), 298. 50) C. Wagner: J. Electrochem. Soc., 65 (1961), 243.

Conclusions

From investigation into the effects of niobium additions and thermomechanical history on austenite grain coarsening in Ti-modified SAE 8620 steel, it was shown that suppression of grain growth in carburizing steels can be achieved with controlled Ti and Nb additions. In addition, the following specific conclusion can be drawn: (1) Solubility considerations in microalloyed steels can be successfully employed to design new steels for potential use in high temperature carburizing processes. (2) Simultaneous control of both Ti and Nb additions to first tie up available nitrogen with Ti and then precipitate Nb as NbC provided an efficient method to produce steels which exhibited controlled austenite grain size during carburizing. (3) To provide optimal resistance to grain growth at higher carburizing temperatures, e.g. up to 1 050°C, Timodified steels with at least approximately 0.06 wt% Nb provide the best potential. The highest levels of Nb soluble during solid state processing (approximately 0.1 wt%) appear to give the finest austenite grain sizes at elevated carburizing temperatures. (4) While secondary to the effects of alloy additions and temperature, thermomechanical processing of carburizing bar steels offers an additional opportunity to process steels for reduced susceptibility to grain growth at carburizing temperatures. Specifically, controlled rolling resulted in the development of distribution of fine precipitates prior to carburizing which were shown to enhance grain growth resistance. Acknowledgments Support is gratefully acknowledged from the sponsors of the Advanced Steel Processing and Products Research Center, an NSF Industry/University Cooperative Research Center at the Colorado School of Mines. Special Thanks is given to the Timken Company for providing the experimental steels critical to this study. One author (AlOgab) gratefully acknowledges the support of the Government of Saudi Arabia through the Ministry of Higher Education. REFERENCES 1) J. I. Goldstein and A. E. Moren: Metall. Trans. A, 9A (1978), 1515. 2) F. E. Harris: Met. Prog., (1943), Aug., 265. 3) H. E. Boyer: Case Hardening of steel, ASM, Metals Park, Ohio, USA (1987), 157. 4) J. P. Wise and D. K. Matlock: SAE Technical Paper Series No. 20001-0611, SAE, Warrendale, PA, (2000). 5) B. E. Cornelissen, D. K. Matlock, G. Krauss, B. Gondesen and F. Hoffmann: SAE Technical Paper No. 1999-01-0602, SAE, Warrendale, PA, (1999). 6) J. L. Pacheco and G. Krauss: Carburizing: Processing and Performance, ed. by G. Krauss, Am. Soc. Met., Metals Park, OH, (1989), 227. 7) S. G. Davidson, J. P. Wise and J. G. Speer: Proc. of the 20th ASM Heat Treating Conf. ASM Int., Metals Park, OH, (2000), 1144. 8) K. C. Evanson, G. Krauss and D. K. Matlock: Grain Growth in Polycrystalline Materials III, ed. by H. Weiland, B. L. Adams and A. D. Rollett, TMS, Warrendale, PA, (1998), 599. 9) Y. Kurebayashi and S. Nakamura: SAE Technical Paper No. 950209, SAE, Warrendale, PA, (1995). 10) W. Bleck, F. Hippenstestiel, F. Hoffmann, C. Pichard, A. Kueper, H. Smith, R. Leppänen and C. Bertrand: European Commission Report, no. EUR 20630, (2003). 11) K. A. Alogab: Colorado School of Mines, Golden, CO, (2004). 12) D. K. Matlock, K. A. Alogab, M. D. Richards and J. G. Speer: Mater.

© 2007 ISIJ

316