Sie sind auf Seite 1von 10

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353.

1999

Microstructural Evolution in a 17-4 PH Stainless Steel after Aging at 400C


M. MURAYAMA, Y. KATAYAMA and K. HONO
The microstructure of 17-4 PH stainless steel at various stages of heat treatment, i.e. after solution heattreatment, tempering at 580C and long term aging at 400C, have been studied by atom probe field ion microscopy (APFIM) and transmission electron microscopy (TEM). The solution treated specimen consists largely of martensite with a small fraction of -ferrite. No precipitates are present in the martensite phase, while spherical fcc-Cu particles are present in the d-ferrite. After tempering for 4 h at 580C, coherent Cu particles precipitate in the martensite phase. At this stage, the Cr concentration in the martensite phase is still uniform. After 5000 h aging at 400C, the martensite spinodaly decomposes into Fe-rich and Cr-enriched . In addition, fine particles of the G-phase (structure type D8a, space group Fm3m ) enriched in Si, Ni and Mn have been found in intimate contact with the Cu precipitates. Following spinodal decomposition of the martensite phase, G-phase precipitation occurs after long-term aging.

I. INTRODUCTION

P RECIPITATION-hardened stainless steels are


widely used as structural materials for chemical and power plants because of their balanced combination of good mechanical properties and adequate corrosion resistance. 17-4 PH stainless steel is a martensitic stainless steel containing approximately 3 wt pct Cu and is strengthened by precipitation of copper in the martensite matrix [1-8]. After a solution heat-treatment, this alloy is precipitation hardened by tempering at about 580C for about 4 hours. Typical service temperatures in power plant applications are below 300C, but increases in hardness and tensile strength accompanied by embrittlement was reported at temperatures ranging from 300 to 400C after long term aging. Since these materials have to serve for a very long period of time during the life span of the plants, understanding the embrittlement mechanism at slightly above the service temperature is very important. The precipitation sequence in 17-4 PH stainless steel begins with formation of coherent copper precipitates, which occurs during the tempering treatment before service. These coherent particles were reported to transform to incoherent fcc-Cu particles after long term aging at temperatures around 400C [3]. In addition, since the Cr concentration in 17-4 PH is within the spinodal line, phase decomposition of the martensite into the Ferich and the Cr-enriched is expected on aging below 450C. Much work has shown that stainless steels are embrittled when phase precipitates by spinodal decomposition [11]. Such embrittlement is anticipated in the 17-4 PH stainless steel as well.

Several studies on the effect of aging 17-4 PH stainless steel were carried out [3-6]. Early work by Anthony [4] proposed mechanical properties of 17-4 PH are influenced by precipitation of phase, but no direct evidence for precipitation was presented. Later, Jack and Kalish [3] observed copper precipitation on aging and correlated it to mechanical property changes; however, there was no mention of phase decomposition in the martensite phase. More recently, Yrieix and Guttmann [10] reported that 17-4 PH stainless steel exhibits high susceptibility to aging embrittlement at 400C, and they concluded that it was essentially due to precipitation. In their study, however, no microstructural observation results were shown. Employing atom probe field ion microscopy (APFIM) and transmission electron microscopy (TEM), Miller and Burke [6] showed direct evidence for a precipitation after aging at 482C. They also reported that significant amounts of iron, nickel and manganese were contained in the -Cu precipitates even in the overaged condition. However, their aging temperature is rather high compared to the service condition of 17-4 PH steel. This study aimed to carry out a more complete characterization of microstructures in 17-4 PH stainless steel at various stages of heat treatment, i.e., after solution heat-treatment, tempering at 580C for four hours, and long term aging at 400C, in order to obtain a better understanding of the embrittlement phenomena on aging. II. EXPERIMENTAL PROCEDURES The chemical composition of the alloy used in this study was Fe-16.5Cr-4.0Ni-3.4Cu-0.6Si-0.6Mn-0.3Nb0.06C (wt pct) or Fe-17.5Cr-3.8Ni-2.9Cu-1.2Si-0.6Mn0.2Nb-0.3C (at. pct). The alloy was solution heat-treated at 1050C for 1 h and subsequently water quenched. The solution treated samples were then aged at 580C for 4 h (tempering). This heat treatment causes precipitation of coherent Cu precipitates in the martensite phase

M. MURAYAMA, Researcher, and K. HONO, Head of 3rd Laboratory, are with Materials Physics Division, National Research Institute for Metals, Tsukuba 305-0047, Japan. Y. KATAYAMA, is with Heavy Apparatus Engineering Laboratory, Toshiba Corporation, Yokohama 230-0045, Japan Manuscript submitted April 21, 1998.

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353. 1999

Table I

Changes in mechanical properties of 17-4 PH by aging


Yield strength (Mpa) Tensile Elonga- Charpy Vnotch energy strength Tion absorp(Mpa) (%) tion (J) 1085 1434 23 8.3 107 3

Pre-aged alloy (580C x 4h) prolonged aged alloy (400C x 5000h)

895 1362

Charpy V-notch energy absorption was measured at 0C.

1600 Yield Strength / MPa 400 C 1400 1200 1000 800 Tempered 10 100 1000 10000

350 C

Aging time / hr

Fig. 1 0.2% yield stress of 17-4 PH stainless steel as a function of aging time at 400C.

tempering treatment, the specimen was aged for 100 and 5000 hours at 400C. For atom probe analyses, a locally built reflectrontype energy compensated time-of-flight atom probe (1DAP) and a three-dimensional atom probe (3DAP) equipped with CAMECAs tomographic atom probe (TAP) detection system [12] were used. One disadvantage of the 3DAP was its poor mass resolution, because it was not equipped with an energy compensator for improving mass resolution. The mass resolution of the 3DAP used in this study was limited to m/m~200 full width at 10 pct maximum (FW10 pct M), which is significantly lower than that obtained using an energy compensated atom probe (~500 FW10 pct M). Thus, Fe2+ and Mn 2+ ions, which have similar mass-to-charge ratios were not distinguished in the 3DAP analyses. Thus, detailed spatial information provided by 3DAP analysis was complemented by 1DAP analysis with a high mass resolution. Field ion microscopy images were observed at temperatures of 30 - 60 K with Ne as an imaging gas, and atom probe analyses were carried out at a specimen temperature of about 30 K, under a UHV (~1x10-10 torr) condition, with a pulse fraction (Vp/Vdc) of 20 % and a pulse repetition rate of 600 Hz. Microstructures of the specimens were examined with a Philips CM200 transmission electron microscope (TEM), operated at 200 kV. High resolution transmission electron microscope (HRTEM) observations were carried out using a JEOL JEM-2000EX, operated at 200 kV. Thin foils for TEM were prepared by grinding the slices to a thickness of about 100 mm, then by twin-jet electropolishing using a 5 pct perchloric acid-acetic acid solution at 287 K. For long-term aged specimen, ion bean thinning was employed for thin foil preparation, because it was found that Cu particles are preferentially dissolved by electropolishing. III. RESULTS

-ferrite NbC

A. Mechanical properties Figure 1 shows the influence of aging times on yield strength at 350 and 400C. For both temperatures, an increase in yield strength occurs after 10 hours aging, and the strengthening response is much faster at 400C. Increase in the yield strength is almost saturated after 10,000 hours aging, and 80 and 90 pct of strengthening is achieved after 100 and 1000 hours aging respectively. Values of yield strength, tensile strength, elongation and Charpy V-notch energy absorption measured before and after 5000 h aging at 400C are summarized in Table I. Increases in yield strength, tensile strength occur after long term aging accompanied by decreases in elongation and Charpy V-notch energy absorption. This indicates that embrittlement occurs as a result of long term aging.

1m

Fig. 2 TEM bright field image of the martensite phase in 17-4 PH stainless steel after solution heat-treatment. The predominant phase is lath martensite. Grains of NbC and -ferrite are indicated.

and provides balanced strength and toughness as shown in Table. I. This is the typical condition of 17-4 PH stainless steel before use as a structural material. After this

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353. 1999

(a)
Fe &Cr / at.%

100 80 60 40 20 0 30

Ni Si Mn

15 0 30 15 0

400nm (b)
011 020fcc

10 0 30

Cu

15 0 0 50 100 150 200 250 300

Number of Atoms / x50

Fig. 3 (a) TEM bright field image of the -ferrite phase in 17-4 PH stainless steel after solution treatment. Fine Cu precipitates are observed. Some of them are apparently associated with dislocations. (b) SAD pattern taken slightly inclined form the [111] zone.

Fig. 4 1DAP concentration depth profiles of Fe, Cr, Ni, Si, Mn and Cu in the martensite phase in 17-4 PH stainless steel after solution treatment.

Table II. EDX analysis results of Cu-rich precipitate in -ferrite, -ferrite and martensite phase in the solution treated condition (all in at. %). Cr Ni Si Cu Mn precipitate 20.9 3.52 0.94 11.8 0.43 -ferrite 22.7 4.00 1.43 3.0 0.40 martensite 17.3 4.93 1.11 3.63 0.34

B. Solution treated microstructure A solution-treated 17-4 PH stainless steel is composed largely of martensite with a minor fraction of dferrite as shown in Figure 2. The martensite phase is consist of a lath structure containing a very high density of dislocations. There is no evidence of precipitates which suggests the martensite phase is supersaturated with Cu and Cr in the solution-treated condition. On the other hand, a high density of fine precipitates are observed in the d-ferrite phase as shown in Figure 3 (a). In the bright-field image, the precipitates

are spherical and each precipitate appears to be associated with dislocation. Absence of strain contrast and presence Moire fringe indicate that the particles are incoherent with the bcc matrix. In fact, selected area diffraction (SAD) pattern taken slightly inclined from the [111] zone show {020}fcc reflections, indicating that the precipitates are fcc-Cu. The orientation relationship (OR) between the particle and the d-ferrite does not match perfectly with the K-S relationship, but has a slight deviation from it. Energy dispersive X-ray (EDX) analysis results shown in Table II indicate that significantly higher Cu concentration is recorded from the precipitate region than from the other regions. This also suggests that the precipitates are Cu, not NbC as previously reported [3]. The EDX results also show that Cr concentration in the -ferrite is slightly higher than that in the martensite phase. Since the -ferrite is a minor constituent phase in this steel, atom probe analyses were carried out only from the martensite phase. Figure 4 shows atom probe concentration depth profiles of the martensite phase in the solution-treated specimen. Horizontal broken lines show average concentration, cav, and statistical errors expected from the number of atoms used for determining local concentrations, cav 2s, where s is the standard deviation. The number of atoms is linearly corre-

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353. 1999

(a)

emental map clearly shows that there is Cu enriched precipitate in this stage. As the SADP (Figure 5(b) and (c)) does not show any evidence for presence of the secondary phase, the Cu enriched precipitate observe in the 3DAP data is believed to be fully coherent bccCu. In order to quantify the concentration of the particle, concentration depth profile was measured from the selected region near the Cu-rich precipitate as shown in Fig. 6 (c). The chemical composition of the Cu-rich precipitates has been found to be 55 at.pct Cu, 30 at.pct Fe, 10 at. pct Cr, 5 at. pct Ni. It is seen that Cr and Ni are rejected from the Cu-enriched particle slightly. It should also be noted that the concentration of Cu in the particle is significantly lower than that expected from the equilibrium e-Cu. D. Aging for 100 h at 400C

100nm
(b) 110 (c) 110

110

011

Fig. 5 (a) TEM bright field image, (b) [001] and (c) [111] SAD pattern of the martensite phase tempered at 580C for 4 h.

lated with the depth scale of the analysis, and the total depth of this analysis is estimated to be approximately 60 nm. This data shows that the martensite in the solution heat-treated specimen is a supersaturated solid solution containing all solute atoms homogeneously. C. Tempered microstructure Figure 5 shows a bright field TEM image of the martensite phase after aging for 4 h at 580C. Brightfield image does not give any clear contrast corresponding to fine coherent Cu precipitates expected in this stage. This suggests that Cu precipitates, if any, is still coherent and they do not cause large strain contrast due to the small strain field around the precipitate. In fact, the SADP taken from a martensite lath does not show any evidence for a secondary phase, suggesting that there is no precipitates with the distinct structure different from the bcc matrix. Figure 6 shows 3DAP elemental mapping of Cu and Cr obtained from the martensite phase. The Cu mapping clearly shows that there is a small spherical particle enriched with Cu. On the other hand, the Cr mapping shows that the distribution of Cr atoms is uniform martensite phase. The Cu el-

Figure 7 (a) and (b) show 3DAP elemental mapping of Cu and Cr obtained from the martensite phase in the specimen aged for 100 h at 400C. The Cu mapping shows that a high density of Cu-rich precipitates of approximately 3 nm in diameter is present. The Cr mapping shows that the distribution of Cr atoms is no longer uniform and fluctuations of Cr concentration occur. Concentration depth profiles of Fe, Cr, Ni, Si and Cu were measured from a selected region cutting two Curich precipitates as shown in Figure 6(c). The concentration of Cu in the Cu-rich precipitates is approximately 70 at. pct Cu, which is much higher than that observed in the tempered specimen, but it is still lower than the equilibrium concentration of the -Cu. The presence of fluctuations in Cr concentration indicates that the phase decomposition occurs in the martensite phase, and it decomposes to Fe-rich a and Cr-enriched phases. The Cr concentration in the phase is only 25 at. pct and this is significantly lower than that of equilibrium phase, suggesting that it is still in an initial stage of the decomposition process. It should be noted that there is no indication of partitioning of Ni. E. Aging for 5000 h at 400C Figure 8 (a) shows a dark field image excited using the 1/2(110) reflection and Figures 8 (b) (d) shows [001], [011], [111] zone SAD patters, respectively obtained from the martensite phase aged for 5000 hours at 400C. Extra reflections indicate an ordered phase having a cube-on-cube orientation relationship precipitates after prolonged aging. The dark field image which was taken using the 1/2(110) reflection shows a high density of fine, ordered particles are dispersed in the martensite matrix. The diffraction pattern was found to be consistent with the G phase (structure type D8a, space group Fm3m ) reported in aged duplex stainless steels [23,26,27] and type 308 stainless steels [19]. Figure 9 shows atom probe concentration depth profiles obtained from the martensite phase in the alloy

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353. 1999

100 Fe & Cr / at. % 80 60 40 20 0 30 Ni 15 0 30 Si 15 0 Cr Fe

(a) Cr
Cu

60 40 20 0 0 2 4 6 8 10

Depth / ~nm

~16nm

(c)

~26nm (b) Cu

~26nm

Fig. 6 3DAP elemental mapping of the martensite phase aged at 580C for 4 h. (a) The uniform distribution of Cr and (b) a fine spherical Cu rich precipitate is observed. (c) Concentration depth profile through a Cu precipitate.

aged for 5000 hours at 400C. In addition to Cu-enriched precipitates, significant fluctuations in Cr concentration are observed. This shows that phase decomposition occurs in the martensite phase, and the martensite decomposes to Fe-rich and Cr-enriched phases. The Cr concentration in the phase is approximately 40 at. pct, which is again significantly lower than that expected from thermal equilibrium (~90 at. pct). Thus, it is believed that this stage is still in the middle of the decomposition process, and has not reached the equilibrium. The Ni concentration profile does not show any tendency of partitioning of Ni to the a phase, unlike previous report by Danoix in the aged ferrite phase of duplex stainless steel [26]. It appears

that Cu enriched precipitates do not have any correlation with the Cr concentration fluctuations. From the concentration profile, two types of Cu precipitates are observed. One is composed of Cu only, and the other is enriched with Ni, Si and Mn, as well. The apparent Cu concentration of the latter is ~20 at. pct Cu. This observation is similar to the solute partitioning in an ultrafine copper-enriched zone in a neutron irradiated A533B submerged arc weld reported by Miller et al [13]. Figure 10 shows 3DAP elemental mappings of Cr, Cu, Ni and Si. Mn cannot be mapped because the mass spectrum of Mn2+ overlaps with that of Fe 2+. The Cr mapping shows the concentration of Cr fluctuates. The Cr enriched regions appear to be interconnected, which

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353. 1999

(a)

(a)

Cr

Cu (b)100
Fe & Cr / at. %
80 60 40 20 0 30

~30nm

~9nm

100nm
Fe Cr
110 200 011 (b) 110 110 (c) 011 (d)

Ni

15 0 30

Fig. 8 TEM (a) dark filed image, (b) [001], (c) [011] and (d) [111] SAD pattern of the martensite phase aged 400C for 5000 h. The dark field image was taken using the 1/2(011) reflection.
100

Si

15 0 70

Fe &Cr / at.%

80 60 40 20 0 30
Ni Si Mn Cu

Cu

50 30 10 0

10

15

20

25

Depth / ~nm

15 0 30 15 0 10

Fig. 7 3DAP elemental mapping of the martensite phase aged at 580C for 4 h. (a) The phase decomposition into Cr enriched and depleted region occur in the martensite phase. (b) Fine spherical Cu rich precipitate. (c) Concentration depth profile obtained from the selected region near the Cu precipitate.

is a typical feature of spinodal decomposition. The periodicity of the fluctuation is on the order of 3 nm. The Cu mapping shows that Cu-enriched particles approximately 8 nm in diameter are present. In direct contact with one of the Cu particle, a Ni and Si enriched particle is observed as indicated by an arrow in Figure 10(b). Such fine particles are always observed in contact with Cu precipitates, and they are believed to be the G-phase based on the TEM observation shown in Figure 8. Figure 10 (c) shows concentration depth profiles across the Cu precipitate and the G-phase indi-

0 30 15 0 0 500 1000 1500 Number of Atoms / x50

Fig. 9 1DAP concentration depth profile of the martensite phase aged at 400C for 5000 h. Ni, Si, and Mn appear to be partitioned into the Cu precipitate.

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353. 1999

Cu precipitate 100 Fe & Cr / at. % 80 60 40 20 0 100 80 60 Ni 40 20 0 50 Cr Fe

G phase

(a) Cr

Si 0 100 80 Cu 60 40 20 0 0 2 4 6 8 10 12 14

Depth / ~nm

~18nm

(c)

~15nm
Ni + Si Cu

~15nm

(b) Cu, Ni and Si


Fig. 10 3DAP elemental mapping of the martensite phase aged at 400C for 5000 h. (a) The phase decomposition into Cr enriched and depleted region occur in the martensite phase. (b) Large dots correspond to Cu atoms and small dots to Ni and Si atoms. (c) Concentration depth profile across the Cu precipitate and the Ni-Si enriched particle obtained from the selected region near the Cu precipitate (indicated by arrow).

cated by the arrow in Figure 10 (b). The Cu precipitate contains approximately 95 at. pct Cu, which is close to the equilibrium concentration of -Cu. The composition of the G-phase is approximately 55 at. pct Ni, 25

at. pct Si, 20 at. pct Fe. In addition, Mn is also enriched in the G-phase based on the 1DAP result shown in Figure 9 (Mn ions can be differentiated from Fe atoms using 1DAP). According to the concentration depth pro-

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353. 1999

(a)
G phase Cu precipitate

the extra reflections near the {110}bcc reflections. The -spacing calculated from the extra reflections is ~0.179 nm which corresponds to the spacing of the {020}fcc planes in fcc-Cu (0.180 nm). The orientation relationship (OR) between the particle and the martensite almost matches with the K-S relationship. IV. DISCUSSION This study has clarified evolution of microstructure in a 17-4 PH stainless steel during long term aging at 400C. Emphasis is given to characterization of chemical features of the precipitates which appear in the martensite phase after prolonged aging. The -Cu precipitates have been found in the -ferrite after the solution heat treatment. The appearance of these precipitates in the -ferrite was unaffected by subsequent aging. Rack and Kalish [3] reported similar microstructural feature in -ferrite, but they attributed them to NbC precipitates. In this study, NbC were observed at martensite lath boundaries with much larger size as shown in Figure 1, and we believe that the fine precipitate in the -ferrite observed in the previous study as well as in this study are the -Cu. Precipitation of Cu in the Fe-Cu binary system has been a subject of numerous studies [14-18], and it is well established that the initial Cu-enriched precipitate is perfectly coherent with the bcc matrix, while large overaged precipitates have an fcc structure with the K-S OR. However, binary alloys were all solution heat-treated in the single phase ferrite region around 900C, while the 17-4 PH stainless steel is solution heat-treated at much higher temperature around 1050C. Diffusivity of Cu in the d-ferrite at this temperature (DCu ~ 1 x 10 -9 cm2/s) is more than three orders of magnitude higher than that in the austenite (DCu ~ 4 x 10-12 cm 2/s), thus we believe that Cu precipitated out from the -ferrite during cooling after the solution heat treatment. Many Cu particles were observed along dislocations, and this would have ease the strain contrast for precipitation of the fcc -Cu. On the other hand, no indication of presence of precipitates are recognized in as-quenched martensite. The solubility of Cu in the austenite phase is up to 7 at. pct at 1100C. However, the diffusivity of Cu in the austenite is orders of magnitude small than that in the ferrite phase. Thus, the precipitation kinetics are much slower in the austenite phase and Cu can be quenched in the martensite phase from the solution heat-treatment temperature. After tempering at 580C for 4 hours, fine Curich precipitates were detected in the martensite phase using 3DAP. The Cu content in the Cu-rich precipitates is significantly lower than the equilibrium value for Cu. They reported that the average copper concentration of small precipitates is approximately 50 at. pct in the earliest stage of the precipitation. The copper concentration in the fine, Cu-rich particles which precipitate during tempering is in good agreement with the early study by Goodman et al. [16]. Unlike in the d-fer-

2nm

(b)
020fcc 011

Fig. 11 (a) HREM image taken at the [111] zone axis of the martensite phase in the alloy aged at 400C for 5000 h. (b) micro-diffraction pattern obtained from the region with the Moire fringe.

file obtained by the conventional atom probe (Figure 9), Ni, Si and Mn atoms appear to be partitioned to the Cu enriched particle indicated by the dashed lines on the left side of the figure. However, this result is probably an artifact caused by the convoluting effect of the probe hole covering both Cu and G-particles because Ni, Si and Mn enrichment is not observed at the other Cu enriched region in Figure 9. In the latter case, the probe hole covered only the Cu particle and missed the G-phase which was adjacent to the Cu particle. These results demonstrates that employment of the 3DAP technique provides more accurate data for characterizing the morphological feature of fine precipitates embedded in the matrix. Figure 11 (a) shows an HREM image taken along the [111] zone axis of the martensite phase in the alloy aged at 400C for 5000 hours. In this image, fringe contrast having a periodicity of two (011)bcc planes is observed. This is consistent with the fringe contrast expected from the G -phase (Ni 16X 6Si 7, X=Fe, Mn, Si, Fm3m , a=0.406 nm). Furthermore, a small particle is observed adjacent to the G-phase, which is believed to be a Cu precipitate. A microdiffraction pattern taken from the region with the Moire fringes is shown in Figure 11 (b). The [111] microdiffraction pattern shows

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353. 1999

rite, these Cu-rich precipitates observed in the martensite phase after tempering are bcc-Cu. The martensite phase decomposes into Fe-rich and Cr-enriched phases by the spinodal mechanism. In addition, fine precipitates of the G-phase has been found in the martensite phase in direct contact with the Cu precipitates after 5000hours aging at 400C. Concentration of Cr in the phase is approximately 25 at. pct after 100 hours, then reach 40 at. pct after 5000 hours aging. This concentration is significantly lower than that expected from the binodal line and is thought to be still in the decomposition process. Since no precipitation of the G-phase was observed after 100 hours aging, the increase in yield strength up to 100 hours is attributed to the spinodal decomposition rather than the formation of the G-phase. G-phase precipitation was reported in type 308 stainless steels [9, 19] and maraging steels [20] as a grain boundary phase. In recent years, it was shown that the G-phase exists in various Fe-Cr-Ni alloys as a dispersed phase in the interior of the grains [21-27]. For example, Auger et al. [27] reported phase separation and precipitation of the G-phase in the ferrite phase of duplex stainless steel. They concluded that the nucleation of the G-phase takes place at the interface between Cr-rich a and Fe-rich a. However, in the case of the 17-4 PH stainless steel, we have found that G-phase precipitation occurs in intimate contact with -Cu precipitates after the decomposition of the martensite have progressed. This result indicates that the Cu precipitates provides heterogeneous nucleation sites for Gphase formation. One reason for this is that the Cu/ martensite interface is a more suitable site for nucleation of the G-phase than the / interface. The dark field image of the long-term aged martensite phase shows the G-phase is dispersed uniformly in the grain rather than grain boundary as commonly observed in other stainless steels. This suggests that the G-phase precipitation itself does not contribute to the embrittlement. In fact, 80 pct of total yield strength increase occurs before precipitation of the G-phase is observed (100 hours aging at 400C). Our atom probe results have shown that the atomic ratio of Ni to Si in the G-phase is 6:4. In addition, Mn and Fe are also contained in the G-phase. The ternary silicide designated as the G-phase is known to be fcc having 116 atoms in a unit cell. The lattice parameter is ~1.12 nm. The structure of this phase is isotypic with Th6Mn23 (structure type D8a, space group Fm3m ), and the ideal composition was proposed as X6Ni16Si7, where X is typically Ti, but can be substituted for other transition element. Considering the qualitative nature of the atom probe result, the determined composition (55 at. pct Ni, 25 at. pct Si, 20 at. pct Fe and some Mn) is reasonably close to the ideal stoichiometry of X6Ni16Si7 when one allows for substitution of Mn and Fe for X [28]. In this study, partitioning of Ni in the phase after

spinodal decomposition of the martensite was not observed. This result is in contrast to the APFIM results by Danoix et al. [26] which reported that nickel is rejected from Cr-enriched phase and partitioned into the Fe-rich phase in the aged ferrite phase of duplex stainless steel. In the equilibrium condition, the solubility limit of Ni in and were estimated to be 4.1 and 0.01 at. pct, respectively, by the ThermoCalc software. Thus, Ni should partition into the phase in 174 PH as well. This suggests Ni partitioning does not occur until decomposition progresses further and the driving force for the partitioning reaction increases. Similar delayed partitioning behavior was reported for Al in an Fe-Cr based alloy [29]. As the Ni concentration of the duplex stainless steel was higher than that in 174 PH, the driving force for Ni partitioning is much higher in the duplex stainless steel. Thus the absence of concurrent partitioning of Ni with spinodal decomposition in 17-4 PH is reasonable. V. CONCLUSIONS The microstructural features of 17-4 PH stainless steel at various stages of heat-treatment has been investigated by APFIM and TEM. The main results are as follows: 1. The fcc Cu-rich particles precipitate in the -ferrite grains during cooling after the solution heat treatment. The precipitates have the K-S orientation relationship with the -ferrite matrix. 2. Tempering at 580C for 4 hours results in precipitation of coherent bcc particles in the martensite phase. The composition of the precipitates is approximately 60 at. pct Cu and Cr, Ni and Si are rejected from the bcc-Cu. The Cr concentration is homogeneous in the martensite phase at this stage. 3. After 100 hours aging at 400C, evidence for spinodal decomposition of the martensite phase into Fe-rich a and Cr-enriched a phase is found. The Cr concentration in the a phase is approximately 25 at. pct, significantly lower than the equilibrium value. The concentration of Cu in the Cu-enriched precipitate is approximately 70 at. pct at this stage. 4. Spinodal decomposition of the martensite phase progresses further after 5000 hours aging. The Cr concentration of the phase is 40 at. pct at this stage. The structure of the coarsened Cu precipitates is fcc and their concentration is almost 100 at. pct Cu. Fine particles of G-phase containing approximately 60 at. pct Ni, 25 at. pct Si and some Fe and Mn are present in direct contact with the Cu precipitates, suggesting that G-phase is heterogeneously nucleated at the martensite/Cu interface. 5. The increase in hardness and yield strength of the 17-4 PH stainless steel after aging at 400C is mostly caused by spinodal decomposition of the martensite phase. G-phase precipitation does not appear to make

Published in Metall. Mater. Trans. A. Vol. 30A, pp. 345-353. 1999

a significant contribution to the embrittlement during aging. ACKNOWLEDGMENTS The authors thank Professor W.T. Reynolds, Jr., Virginia Polytechnic Institute and State University, for valuable discussions. This work was supported by the Frontier Research Center for Structural Materials, NRIM. REFERENCES
1. G.N. Goller and W.C. Clarke, Jr.: Iron Age, 1950, vol. 165, pp. 79-83. 2. G.N. Goller and W.C. Clarke, Jr.: Iron Age, 1950, vol. 165, pp. 86-89. 3. H.J. Rack and D. Kalish: Metall. Trans., 1974, vol. 5, pp. 15951605. 4. K.C. Antony: J. Metall., 1963, vol. 15, pp. 922-927. 5. U.K. Viswanathan, S. Banerjee and R. Krishnan: Mater. Sci. Eng., 1988, vol. A104, pp. 181-189. 6. M.K. Miller and M.G. Burke: Proc. 5th Int. Smp. on the Environmental Degradation of Materials in Nuclear Power Systems - Water Reactors, September 1991 Monterey, ed. E.P. Simonen , American Nuclear Society, La Grange Park, 1992, pp.689-695. 7. H.R. Habibi-Bajguirani and M.L. Jenkins: Phil. Mag. Lett ., 1966, vol. 73, pp. 155-162. 8. K. Ozbaysal and O.T. Inal: Mater. Sci. Eng., 1990, vol. A130, pp. 205-217. 9. J.M. Vitek, S.A. David, D.J. Alexander, J.R. Keiser and R.K. Nanstad: Acta Metall., 1991, vol. 39, pp. 503-516. 10. B. Yrieix and M. Guttmann: Mater. Sci. Technol, 1993, vol. 9, pp. 125-134.

11. W.C. Leslie: The Physical Metallurgy of Steels, McGraw-Hill Inc., New York, 1981. 12. D. Blavette, B. Deconihout, A. Bostel, J.M. Sarrau, M. Bouet and A. Menand: Rev. Sci. Instrum, 1993, vol. 64, pp. 29112919. 13. M. K. Miller, M. G. Hetherington and M. G. Burke: Metall. Trans. A, 1989, vol. 20A, pp. 2651-2661. 14. P.J. Othen, M.L. Jenkins and G.D.W. Smith: Phil. Mag. A, 1996, vol. 73, pp. 1-24. 15. E. Hornbogen and R.C. Glenn: Trans. AIME, 1960, vol. 218, pp. 1064-1070. 16. S.R. Goodman, S.S. Brenner and J.R. Low, Jr.: Metall. Trans., 1973, vol. 4, pp. 2363-2369. 17. S.R. Goodman, S.S. Brenner and J.R. Low, Jr.: Metall. Trans., 1973, vol. 4, pp. 2371-2378. 18. W.J. Phythian and C.A. English: J. Nucl. Mater., 1993, vol. 205, pp. 162-177. 19. J.M. Vitek: Metall. Trans. A, 1987, vol. 18A, pp. 154-156. 20. H.J. Rack and D. Kalish: Metall. Trans., 1971, vol. 2, pp. 30113020. 21. W. Sha, A. Cerezo and G.D.W. Smith: Metall. Trans. A, 1993, vol. 24A, pp. 1241-1249. 22. R.C. Ecob, R.C. Lobb and V.L. Kohler: J. Mater. Sci., 1987, vol. 22, pp. 2867-2880. 23. M. Vrinat, R. Cozar and Y. Meyzaud: Scripta Metall ., 1986, vol. 20, pp. 1101-1106. 24. B. Soylu and R.W.K. Honeycombe: Mater. Sci. Technol., 1991, vol. 7, pp. 137-145. 25. H.M. Chung: Int. J. Pres. Ves. Piping, 1992, vol. 50, pp. 179213. 26. F. Danoix, P. Auger and D. Blavette: Surf. Sci., 1992, vol. 266, pp. 364-369. 27. P. Auger, F. Danoix, A. Menand, S. Bonnet, J. Bourgoin and M. Guttmann: Mater. Sci. Technol., 1990, vol. 6, pp. 301-313. 28. F.X. Spiegel, D. Bardos and P.A. Beck: Trans. AIME , 1963, vol. 227, pp. 575-579. 29. H.G. Read, H. Murakami and K. Hono: Scripta Mater., 1997, vol. 36, pp. 355-361.

Das könnte Ihnen auch gefallen