Sei sulla pagina 1di 11

Journal of Nuclear Materials 439 (2013) 148158

Contents lists available at SciVerse ScienceDirect

Journal of Nuclear Materials


journal homepage: www.elsevier.com/locate/jnucmat

Microstructural defect evolution in neutron Irradiated 12Cr18Ni9Ti stainless steel during subsequent isochronous annealing
K.V. Tsay a, O.P. Maksimkin a, L.G. Turubarova a, O.V. Rofman a, F.A. Garner b,
a b

Institute of Nuclear Physics NNC RK, Almaty, Kazakhstan Radiation Effects Consulting, Richland, WA, USA

a r t i c l e

i n f o

a b s t r a c t
Transmission electron microscopy and microhardness measurements were used to examine changes in microstructure and associated strengthening induced in austenitic stainless steel 12Cr18Ni9Ti irradiated to 0.001 and 5 dpa in the WWR-K reactor before and after being subjected to post-irradiation isochronal annealing. The relatively low values of irradiation temperature and dpa rate (80 C and 1.2 108 dpa/s) experienced by this steel allowed characterization of defect microstructures over a wide range of defect ensembles, all at constant composition, produced rst by irradiation and then by annealing at temperatures between 450 and 1050 C. It was shown that the dispersed barrier hardening model with commonly accepted physical properties successfully predicted the observed hardening. It was also observed that when TiC precipitates form at higher annealing temperatures, the alloy does not change in hardness, reecting a balance between precipitate-hardening and matrix-softening due to removal of solute-strengthening elements titanium and carbon. Such matrix-softening is not often considered in other studies, especially where the contribution of precipitates to hardening is a secondorder effect. 2013 Elsevier B.V. All rights reserved.

Article history: Received 29 July 2011 Accepted 29 March 2013 Available online 10 April 2013

1. Introduction It is well known that the conditions of neutron irradiation, especially dose, dose rate and irradiation temperature, lead to signicant changes in both microstructure and macroscopic properties of irradiated stainless steels [121]. Much of the published experimental data accumulated to date focuses on the behavior of steels irradiated at relatively high dpa rates and high irradiation temperatures. In fast reactors these temperatures are usually in the range 320700 C, and in light water cooled reactors are in the range 280380 C. Many small test reactors, however, have low inlet temperatures in the range 25100 C. The WWR-K test reactor, located in Almaty, Kazakhstan is an example of the latter type of reactor with an inlet temperature of 80 C. While some data are available on radiation-induced changes in austenitic steels following long-term irradiation at temperatures below 300 C and at relatively low dpa rates, the total data base is rather sparse, especially for austenitic steels used to construct reactors in the states of the former Soviet Union. Investigation of steels irradiated in research reactors such as WWR-K can be used to extend our understanding of radiation-induced microstructural evolution and hardening of such steels to lower temperatures.

In this paper we examine the austenitic stainless steel 12Cr18Ni9Ti (a Russian analog of AISI 321 SS) irradiated at low temperature and low dpa rates. In order to extend the scientic value of this investigation we used isochronal annealing over a large temperature range to obtain a wide range of defect ensembles, all at identical composition, thereby enabling us to better test the generality of the dispersed barrier hardening model when coupled with microscopy observation. The predictions of microstructurebased modeling calculations are then compared with microhardness measurements. 2. Experimental procedure The component being investigated served as the central core of an automatic control rod used to regulate the neutron ux and power in WWR-K, a light water-cooled reactor which operated at 6 MW thermal power. See Fig. 1a for details of rod construction and size. This control rod was clad with an aluminum alloy designated SAV-1 and results were presented previously that focused on the microstructural evolution of the aluminum alloy [22]. The absorber material of the control rod was Russian austenitic steel 12Cr18Ni9Ti, whose chemical composition is presented in Table 1. Note that in the Russian alloy labeling system, 0.12 wt.% carbon is the maximum allowable in 12Cr18Ni9Ti, but actual commercial steels usually contain some level of carbon below the maximum. The actual carbon was not measured in this

Corresponding author. Tel.: +1 509 521 1633.


E-mail address: frank.garner@dslextreme.com (F.A. Garner). 0022-3115/$ - see front matter 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.jnucmat.2013.03.086

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158

149

Fig. 1. (a) Schematic image of the control rod; (b) cutting scheme employed for preparing specimens for investigation.

Table 1 Specication of chemical composition in wt.% of 12Cr18Ni9Ti stainless steel. C 0.12 Cr 17.0 Ni 9.5 Ti 0.6 Si 0.34 Mn 1.6 S 0.01 P 0.02 Cu 0.2

material and no archive material was available, issues that will be discussed later in the paper. Prior to construction and irradiation the steel was thermomechanically treated, involving cold-rolling with 15% reduction and subsequent annealing at 800 C for 1 h. The bottom end of the steel core was maintained at 80 C for 20 years, experiencing an average dpa rate of 1.2 108 dpa/s, reaching a total damage exposure of 5 dpa calculated using the standard NRT model. The top end of the steel core was located outside of the reactor core and reached a much lower exposure, estimated to be 6 0.001 dpa. The temperatures of the two sections are estimated to be within 5 C of each other, with gamma heating elevating the lower segment above the 80 C coolant temperature. Hereafter we will refer to the temperature as 80 C. Right-circular segments of 4 mm thickness were cut from both the bottom and top ends of the steel core to compare the radiationinduced microstructures developed at 5 dpa and 60.001, but most attention was focused on the 5 dpa section. The segments were then sliced to produce parallel-sided plates of 18 4 0.3 mm, as shown in Fig. 1b. Finally, 3 mm diameter disks for transmission electron microscopy were punched from the plates. Thinning of the disks was accomplished by electro-polishing at 16 C in a solution of 20 ml HClO4 + 70 ml C2H5OH + 70 ml C3H8O. A set of one hour isochronous annealing treatments were performed on the disks in a vacuum furnace (103 Torr), one at 200 C, and others with 100 C steps over the range 4501050 C. As shown in Fig. 1 examinations were performed on disks removed from near the core axis (<2 mm from the rod axis) and from the near-surface periphery region of the rod to determine whether the damage changed with depth from the rod surface. Grain sizes were estimated on irradiated specimens using the NEOPHOT-2 metallography microscope. Vickers microhardness measurements using a 50 g load on the PMT-3 hardness device were performed on the disks after irradiation and after each subsequent annealing condition. The microstructure was studied primarily at specimens irradiated to 5 dpa and annealed at temperatures 4501050 C using standard techniques on a JEM100CX transmission electron microscope operating at 100 kV. 3. Results and discussion 3.1. Characterization of the 12Cr18Ni9Ti steel after irradiation At the top of the control rod core a rather nonhomogeneous banded microstructure was observed, varying somewhat from

the center to the periphery of the rod. The microstructure was characterized by the presence of wide work-hardened bands (width 300 lm) with a mean grain size of 10 lm, as well as irregularly distributed layers (or bands with width <50 lm) of larger grains with an average size of 17 lm. Average values of microhardness in these bands were equal to 244 and 200 kg/mm2, respectively. Comparison of microhardness value in zones with larger grains to that in non-irradiated steel after 1 h-austenitizing annealing at 1050 C (199 kg/mm2) indicates negligible in-reactor hardening of the material. It follows from this, that hardening in wide work-hardened bands arose from initial mechanicalthermal treatment of the rod material, and not from irradiation. Therefore, we will assume later that the material at the top of the rod was essentially unirradiated, and therefore can serve as a reference state in the absence of unirradiated archive material. Steel from the bottom of the rod exhibited a homogeneous microstructure from center to periphery, without deformation bands, but did possess a grain texture, with grains from 10 to 40 lm in size oriented parallel to the rod axis. The average microhardness at 5 dpa was 390 kg/mm2, demonstrating that signicant radiation hardening had occurred. Microscopy studies conrmed that modication of microstructure occurs during irradiation to 5 dpa (Fig. 2). The microstructure in the top segment did not differ much from that of non-irradiated steel (Fig. 2a), and was characterized by network dislocations (1.5 1010 cm2) and stacking faults. The microstructure of the bottom segment at 5 dpa contained a high density of Frank loops and small black dot clusters (Fig. 2b). The black dots are imaged as white dots in the dark-eld micrograph. Frank loops are faulted dislocation loops with a/3h1 1 1i Burgers vector, originating from coalescence of either interstitials or vacancies on {1 1 1} planes. At g = 002 stacking faults were easily observed inside large loops with sizes of 15 nm and larger. The average density of Frank loops was 6.6 1021 m3, with loop sizes ranging from 5 to 60 nm with a peak at 15 nm and average size of 23 nm. The majority of Frank loops with sizes of P15 nm were of interstitial type as conrmed using classical Inside/Outside imaging techniques [23]. Large Frank loops were arrayed in chains giving the appearance of a labyrinth network structure. Segments of the network appear to be quasi-cells elongated along the (0 0 2) direction with a cell size of 0.1 lm, corresponding approximately to the thickness of the microscopy foil area under examination. At the same time black dot defects, with an average size of 2.3 nm, were uniformly distributed inside the network segments at high density, 1.4 1023 m3. Note that black dots are referred to as defects with size less than 5 nm and mean size of 1.5 nm [25]. In Ref. [2] the maximum size for such defects was limited to 34 nm. In comparison, 316LN steel irradiated to 3 dpa at 90 C exhibited a mean size of black dots at 1.2 nm [26].

150

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158

(a)

(b)

100nm

100nm

Fig. 2. Micrographs of 12Cr18Ni9Ti steel irradiated at 0.001 dpa (a) and 5 dpa (b). Micrograph of 5 dpa-irradiated steel was imaged in dark eld using the 0 0 2 reection.

The nature of black dots was not examined in this work and these defects could be either vacancy or interstitial in character. However, it is necessary to point out that most referenced data were produced at higher dpa rates of neutron irradiation in commercial ssion reactors in comparison with the very low dpa rate of the WWR-K research reactor. Other research shows that small vacancy defects also prevail in steels irradiated at similar low temperature conditions. For instance, several studies provide evidence that in a Fe16Ni15Cr alloy irradiated with neutrons at 80 C, the number of vacancy clusters exceeded the number of interstitial clusters [24,25]. Our obtained values of densities for Frank loops and black dots formed in 12Cr18Ni9Ti steel at 5 dpa and 80 C are similar to the data for SA316 and SA347 steels at 17 dpa and 55 C [2], and for 316LN at 3 dpa and 90 C [26]. The peculiarity in our case is the larger sizes of both defect types, perhaps caused by the longer exposure time during neutron irradiation at substantially lower dpa rates. 3.2. Changes in microstructure of 5 dpa irradiated steel after isochronous annealing in the range 4501050 C With one exception, metallographic examinations conducted on the 5 dpa specimens after isochronous annealing showed no noticeable change in average size of grains or in secondary inclusions. Annealing at 1050 C produced the only exception, with strong growth occurring in grains and accompanied by growth in size and density of coarse carbide particles that formed both in the matrix and on grain boundaries. Microscopy investigations conducted after isochronous annealing revealed evolution in the populations of dislocation loops, small defects, nely-dispersed secondary precipitates of titanium carbides, globular TiC particles and plate-shaped precipitates of M23C6 type, as well as development of a low density of helium bubbles. The stages of temperature-induced transformation in microstructure after annealing at different temperatures are illustrated by micrographs in Figs. 3 and 4. Values of density and size for loops and black dots are given in Table 2 and the size distributions of loops are presented in Fig. 5. 3.2.1. Evolution of Frank loop, black dot defects and dislocation networks in 5 dpa steel After 1 h annealing at 450 C there was a doubling of dislocation loop density and a noticeable decrease of the average size by a factor of 1.4. The peak of Frank loop size distribution shifted to lower sizes and the density of ne-scale black dots decreased more than fourfold. After annealing at 550 C, the defect microstructure consisted of a mixture of faulted Frank loops, perfect loops and black dots that were homogeneously distributed in-between dislocation loops.

The loop density had decreased slightly in comparison with annealing at 450 C, whereas the average size changed only negligibly. Perfect loops with a Burgers vector a=2h1 1 0i are known to develop from interaction between faulted Frank loops and Shock i ! a=2h1 1 0i ley partial dislocations via the a=3h1 1 1i a=6h1 1 2 reaction. Dark-eld micrographs using the 0 0 2 reection reveal both faulted Frank loops and perfect loops without stacking faults. The fraction of Frank loops was around 30% of the total loop population after annealing at 550 C. After annealing at 650 C, no black dot defects remained. Only a few residual Frank loops, about 10% of the original loop population, were observed as shown in Fig. 5c. The total loop content decreased by half in comparison with annealing at 550 C, and the average size grew by a factor of 1.7. Some loops with sizes as large as 90 nm had formed. Values of loop density and size-distribution were similar to the Frank loop characteristics observed before annealing. Following annealing at 750 C, the average loop size increased by a factor of 1.5 and the density decreased more than by an order of magnitude as compared with measurements made before thermal treatment. The loop network had disappeared and growing loops had formed tangled dislocation networks. The density of dislocations increased as the loop density decreased. After annealing at 800 C, there were no noticeable changes compared with that observed after annealing at 750 C. After annealing at 850 C, grains in the periphery specimens contained only dislocation networks. Loops were occasionally observed in specimens taken from near the core axis. Sizes of some of these dislocation loops reached 120 nm. After annealing at higher temperatures of 9501050 C no dislocation loops were observed. 3.2.2. Thermally-induced precipitates After annealing at 750 C, a great number of nely dispersed precipitates were found on loops and dislocations (Fig. 6). The precipitates were partly coherent with the matrix and their images were similar in contrast to those of perfect dislocation loops when imaged in bright-eld micrographs. The similarity between diffraction patterns obtained from these precipitates and the austenitic matrix at various annealing temperatures conrms a cube-to-cube orientation relationship. The estimated lattice parameter for the precipitates is 0.420 nm, close to the lattice parameter of fcc TiC at 0.42 to 0.433 nm or possibly titanium carbo-nitrides Ti(C,N) [27]. Initially there is no measurable nitrogen in the steel; however, some nitrogen may have been introduced into the material during annealing. The lattice parameter of the austenitic matrix, 0.359 nm, was obtained using X-ray diffractometry. Table 3 shows the maximum observed density of nely dispersed precipitates observed after annealing at 750 C, a

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158

151

(a)

(b)

(c)

(d)

(e)

(f)

100 nm

Fig. 3. Temperature-induced evolution of dislocation loops as a result of post-irradiation isochronous annealing: (a) not annealed; (b) 450 C; (c) 550 C; (d) 650 C; (e) 750 C; (f) 850 C.

temperature where the density of interstitial dislocation loops had decreased strongly. The density of precipitates was observed to decrease strongly as the annealing temperature increased. In the interval from 750 to 850 C the average size of precipitates slightly decreased and then, with increasing temperature (9501050 C), again began to increase. After annealing at 1050 C such precipitates were only rarely seen. It was noted that, while the average size increased by a factor of 2, the precipitates lost coherence with the matrix and became non-transparent to the electron beam. A possible mechanism for the formation of the nely-dispersed precipitates in irradiated steel during heat treatment may be accumulation of impurity atoms at faulted Frank loops by the Suzuki mechanism. It is known that solubility of impurity atoms in fcc metals at faulted Frank loops in average is higher than that in the matrix. Annealing is likely to result in acceleration of impurity diffusion towards edge dislocations of perfect loops. Flow of impurity atoms towards dislocation loops appears to lead to intensive formation of ne particles. At the same time the rate of particle formation is dependent on transformations of faulted Frank loops into perfect loops. This phenomenon was observed in 0Cr16Ni15Mo3Nb steel doped with He to concentrations of 5 105 to 0.1 at.% after heat treatment at temperatures of 500 750 C [28,29]. The annealing temperature of 550 C corresponds to the point where fault-free dislocation loops begin to form in the microstruc-

ture. Despite observations showing the presence of thickening on perfect dislocation loops (Fig. 7b), electron diffraction does not yet show additional reections associated with the presence of a second phase. It could be assumed that the number of formed lattice planes in the second phase particles is quite small, and that the observed objects are GuinierPreston type segregation zones. Annealing at P750 C produceses perfect loops heavily decorated with precipitates of MC type (Fig. 7c) which give clear micro-diffraction reections. As the density of precipitates reaches its maximum for the studied range of temperatures, then it is likely that segregation zones start to transform into precipitates at temperatures slightly below 750 C. 3.2.3. Effect of thermal treatment on microstructural homogeneity It is known that the physicalmechanical properties of a material are sometimes dominated by grain boundary effects. Grain boundaries represent strong sinks in competition with loops and clusters and strongly emit vacancies at higher temperatures that decrease the survivability of nearby defect clusters, especially interstitial clusters and loops. The presence of sufciently wide zones depleted in loop and cluster defects located next to grain boundaries can cause them to be areas of minimal resistance during deformation processes. Microscopy examination showed that at annealing temperatures P550 C the width of defect depleted zones increased up to 800 C (Fig. 8). For temperatures P800 C

152

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158

(a)

(b)

100 nm

(c)

Fig. 4. Reduction of black dots and faulted dislocation loops with increasing annealing temperature, as shown in dark-eld micrographs after isochronous annealing at (a) 450 C (0 0 2 reection), (b) 550 C (0 0 2 reection) and at (c) 650 C (1 1 1) reection.

Table 2 Summary of dislocation loop and black dot defect density and sizes in irradiated (5 dpa) and thermally treated 12Cr18Ni9Ti steel (The data from central and peripheral specimens were averaged.). Dislocation loops Annealing temperature (C) Not annealed 450 550 650 750 800 850 Density (m3) 6.6 10 1.3 1022 1.2 1022 5.7 1021 6.3 1020 3.9 1020
21

Black dots Average size (nm) 23.0 16.7 17.2 29.0 42.8 55.6 Size range (nm) 560 550 550 1090 15100 2590 30120 Density (m3) 1.4 1023 3.1 1022 1.2 1022 Average size (nm) 2.3

the width of depleted zones became comparable to the grain size. At 800 C this effect is pronounced for small grains, but at higher annealing temperatures, when the spread in sizes for different grains decreased, it became pronounced for grains with sizes of several tens of microns. After annealing at 1050 C the steel microstructure is close to that of the original austenitic condition. 3.2.4. Helium bubbles TEM showed that annealing temperatures P800 C resulted in nucleation of helium bubbles in grain boundary zones that were depleted of other defects (Table 4). In the matrix helium bubbles were observed to have nucleated predominantly on dislocations. With increasing temperature the average size of bubbles grew and their density decreased. For instance, after annealing at 1050 C, the density of helium bubbles decreased by a factor of 2.6 and their average size increased by a factor of 3.5 as compared with annealing at 800 C. 3.2.5. Globular and plate-shaped carbide precipitates After annealing at 1050 C an almost complete dissolution of nely dispersed TiC precipitates was observed, as well as formation

of two types of new precipitates. The rst new precipitate type was granular particles of TiC which were partially coherent with the matrix and formed on dislocations, grain boundaries and twin boundaries (Fig. 9a). The second type of new precipitates was plate-shape precipitates growing on grain boundaries (Fig. 9b). These precipitates were partially coherent with the matrix, and had a parallel orientation relationship corresponding to a cubic lattice

  101 c jjf101gphase ; h111ic jjh111iphase


with reections approximately 1/3 of the matrix reections. The estimated lattice parameter a = 1.067 nm corresponds well with M23C6 carbide (a = 1.0571.068 nm) [27,30]. 3.3. Microhardness change after irradiation and isochronous annealing The results of microhardness measurement obtained on specimens from the bottom of the steel rod are presented in Fig. 10. These specimens were irradiated to 5 dpa and then post-irradiation annealed. Note that the experimental point on the graph at

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158

153

Fig. 5. Loop size distributions prior to and after isochronous annealing.

(a)

(b)

100 nm

100 nm

Fig. 6. Finely-dispersed precipitates after annealing: (a) at 750 C, 1 1 0c zone axis for the austenitic matrix; (b) at 950 C, 1 1 2c .

Table 3 Characteristics of nely-dispersed precipitates (The data from central and peripheral specimens were averaged.). Annealing temperature (C) 750 800 850 950 1050 Regions with precipitates (%) 100 85 70 55 Rarely observed Average density (m3) 1.0 1022 6.7 1021 5.8 1021 3.2 1021 Average size (nm) 16.3 14.3 13.5 15.8 Size interval (nm) 1025 1025 1040 1040 1050

80 C corresponds to irradiated material without annealing. As the dpa at this level is very low, then it is reasonable to consider this material as annealed at the given temperature and present these data on the same graph with the results of annealing for comparison.

Analysis of microhardness changes with annealing temperature in irradiated steel has revealed a signicant decrease in radiationinduced hardening in the interval of 450650 C. After annealing at 650 C, quantitative characteristics of dislocation loops in the material were similar to those of the material before annealing.

154

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158

(a)

(b)

(c)

35 nm

35 nm

35 nm

Fig. 7. Formation of nely dispersed precipitates on dislocation loops in 12Cr18Ni9Ti steel: (a) after irradiation, (b) after annealing at 550 C, (c) after annealing at 750 C.

Fig. 8. Zones depleted of loops and clusters observed after various treatments. (a) without annealing, (b) annealed at 650 C, (c) annealed at 800 C, (d) average width of defect-depleted zones versus annealing temperature.

Table 4 Average characteristics of helium bubbles. Annealing Regions with temperature (C) bubbles (%) 800 850 950 1050 Average Average Size density (m3) size (nm) interval (nm) 8.8 13.5 28.6 515 520 1025 1075

Rarely observed 15 3.7 1020 45 3.2 1020 100 1.4 1020

However, Frank loops were transformed into perfect loops and black dot clusters fully disappeared from the microstructure. This led to a 70% loss of radiation-induced increase in microhardness. Therefore, the critical role of dislocation loops in radiation-induced hardening of 12Cr18Ni9Ti steel suggests that their density is rather less important than the type of dislocation loops and the presence of an initially high concentration of black dots. It is also important to note that formation of numerous nely-dispersed precipitates in the range 750950 C did not lead to the expected additional hardening. Experimental values of irradiation hardening Hl are determined from the difference between the microhardness of irradiated and

initial states. In this work essentially unirradiated material from the top end of the rod was assumed to be the initial state of the steel before the irradiation and annealing procedures. This material (Fig. 10) shows some hardness increment in comparison with the austenized steel, most likely being a consequence of deformation hardening during fabrication. The microhardness value of the initial state of the steel rod, Hl0, was accepted equal to 244 kg/ mm2, appropriate to the predominant microstructure of workhardened bands. Annealing inuenced both the microstructure and hardness properties of the irradiated steel, leading to transformation of radiation-induced defects, as well as to appearance of new radiationinduced and thermally-induced defects. As a result, the hardening effect after annealing is determined by both the radiation and thermal history. A major issue is how to specify the initial state for the material annealed after irradiation. It is assumed that the original material should be identical to that of the rod prior to irradiation and it should be annealed at the same temperature. For this purpose annealing for 1 h at 450950 C was carried out for specimens from the top end of the rod. Microhardness measurements were also performed. It was shown that annealing at temperatures

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158

155

(a)

(b)

120 nm

200nm

Fig. 9. Microstructure after annealing at 1050 C: (a) helium bubbles and globular precipitates of TiC; (b) plate-shaped M23C6 precipitates coating grain boundaries.

after irradiaton 400

, kg/mm

350

exp. data, bottom - center exp. data, bottom - periphery exp. data, top after irrad. calc. with FL=0.5 after irrad. calc. with FL=0.33 after annealing calc.

300

250

200

150

300

450

600

750

900

1050 1200

Temperature of Isochronal Annealing,oC


Fig. 10. Microhardness of 12X18H9T steel from the bottom and the top ends of the control rod after irradiation and post-irradiation annealing. Also shown is a comparison of measured data with microstructure-based predictions at 5 dpa.

P450 C resulted in disappearance of deformation-induced hardening in the work-hardened zones. In so doing, Hl decreased to the value of 199 kg/mm2 (austenized steel) and did not change with further temperature rise. This means that after annealing at temperatures 4501050 C Hl of austenized steel could be taken as Hl0 of the initial material. 3.4. Discussion The relatively low irradiation temperature in the WWR-K reactor leads the authors to assume that interstitial atoms are predominantly involved in diffusional processes, interacting with microstructural sinks, while vacancy migration is limited, probably resulting in formation of vacancy clusters directly at displacement cascade sites. As a result, the major evolution involves interstitialtype defects. This assumption appears to be conrmed by the observed presence of a well-developed spatial network of large Frank loops, mainly of interstitial type. Additionally, this interpretation is supported by the observation that the internal regions of the labyrinth interstitial loop network are homogeneously lled with small defects and black dots presumed to be of both vacancy and interstitial nature. Note that the small loop size distribution after irradiation is similar to those observed in materials irradiated at temperatures P300 C and at higher damage dose rates, or that have undergone post-irradiation heat treatment [12,31]. Temperature increases during annealing activate migration of point defects towards sinks, inducing point defect recombination, dissolution of small defects, as well as growth of large defects. In austenite steels vacancy clusters become thermally unstable, and break up by means of vacancy emission at annealing temperatures

>0.2Tmelt [32,33]. Emission decay of interstitial defects is triggered at higher temperatures (0.30.4Tmelt) [34]. Therefore, annealing temperatures of 450550 C and higher meet the conditions needed to activate emission of both vacancies and interstitial atoms. After isochronous annealing at 450 C, the density of Frank dislocation loops of interstitial type increased by almost a factor of two, and the mean loop size decreased 1.4 times compared to that of non-annealed material. At the same time, a number of black dot clusters decreased by a factor of two. Accordingly [12] some of black dots must have the nature of very small Frank loops. At 450550 C, the temperature is high enough to cause emission decay of small interstitial clusters, which are often not seen in TEM due to their small size. At the same time, somewhat larger clusters, thermodynamically more stable at a given annealing temperature, grow via vacancy emission at the expense of capture of migrating interstitial atoms. As a result, these clusters may reach sizes large enough to identify them by microscopy. In internal segments of the spatial lattice formed by chains of interstitial loops there is a higher density of mobile vacancies at increasingly higher temperatures. The probability for vacancy recombination with neighboring interstitial defects increases with their increasing concentration. A higher intensity of recombination processes during annealing could be associated with the signicant decrease of large loops. Loop sizes P50 nm are disappearing, while the percentages of small and medium loop sizes are signicantly increased. A subject of frequent investigation is the origin of radiation-induced changes in hardness, strengthening and mechanical properties arising from microstructural alteration [26,3439]. In some cases, other studies have also used post-irradiation annealing of 304 and 316 stainless steels to investigate the relationship of microstructure to hardening, elemental segregation and corrosion [4044]. It is worth noting, however, that these two 300 series steels do not contain titanium, and their precipitation behavior was not as pronounced during short-time annealing as it was in 12Cr18Ni9Ti. According to the model of dispersed barrier hardening [45,46] an increase in yield strength DrY is related to the total force needed for a moving dislocation to overcome stress elds around obstacles of power a. The change in yield stress is dened as:

p Dry Malb Nd;

where the Taylor factor M = 3.06 relates the shear stress on a slip plane in a single crystal to the applied tensile stress necessary to activate slip in a polycrystal [47], l is the shear modulus of the matrix taken from [17], b is the Burgers vector of a moving dislocation, and N and d, correspondingly, p are the density and the average diameter of defects. Here 1= Nd denes the average spacing for rigid non-shearable obstacles.

156

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158

This model usually provides a sufcient description of hardening in the case of steels irradiated at temperatures >300 C when the microstructure has many easily observable and measureable voids, precipitates and dislocations, but due to high temperatures a very low density or absence of cluster-type defects. The contributions from different type barriers are combined via superposition [45] as follows:

DrY

r X DrSR;i 2
i

where DrSR;i is the change due to i-th short range obstacle. In order to convert predicted changes in yield stress to predicted microhardness changes the hardness-yield strength correlation [4850]

DrY 3:03DHl ;

was employed where Hl is in kg/mm2 and DrY is in MPa. The parameters used in formulae (1)(3) for calculation of hardening in irradiated to 5 dpa and annealed specimens are given in Table 5. The density of Frank loops and perfect loops is taken as the product of overall density of loops and weighting coefcients for each type of defect. The calculated values of Hl (dened as the sum of Hl0 + Hl) at different annealing temperatures in the interval 4501050 C are in good agreement with obtained experimental data (see Fig. 10). Note that values of measured hardening are somewhat different for the central and periphery regions at temperatures P450 C. Such variation is not necessarily surprising since swaged rods usually possess non-uniform distributions of cold work and sometimes minor compositional variations that develop during cooling. Since the microstructural densities presented in this paper are averaged from the two regions there is some inuence on the predictions of hardness, but these variations are probably of second-order importance compared to the observed inhomogeneities in defect distributions, especially when observed at higher annealing temperatures. As follows from Fig. 8d the width of defect-depleted zones at P800 C is getting comparable to the mean grain size and forms a non-hardened zone which allows easy sliding of dislocations. This issue cannot be taken into account in the models of hardening which assume the presence of uniform defect microstructure in which the width of near-boundary defect-free zones is signicantly lower than the mean grain size. With one exception, the values of a coefcients used in this study for the different defects in the range 4501050 C (see Table 5) agree rather well with the values obtained from other studies performed at irradiation temperatures >300 C. However, the value aP = 0.2 for precipitates used in our calculation is lower compared with the values usually found in the literature where aP = 0.330.45 [38]. There are some possible explanations for this difference. First of all, second phases growing at a rapid rate during postirradiation annealing, often lose their coherence with matrix and perhaps lose some of their hardening ability. More importantly, Eqs. (1) and (2) are usually applied to materials where the precipitates are not the dominant hardening defect, and whose formation

Table 5 Constants for calculation of yield strength increment of 12Cr18Ni9Ti steel after irradiation and annealing. Defects Frank loops Perfect loops, black dot defects Finely-dispersed precipitates of MCtype Small helium bubbles

a
0.33 0.2 0.2 0.2

l
7.87 1010 Pa

b 2.5 1010 m

or dissolution does not substantially alter the matrix composition. In earlier studies no mention is usually made of the consequences of solute removal from the matrix. If precipitation removes from the matrix those elements that are particularly efcient as solution-hardening contributions, the matrix must soften, a consideration not reected in Eqs. (1) and (2). Thus, precipitation can have both positive and negative consequences on hardening, and ignoring the matrix-softening will lead to different values of aP, especially in alloys containing signicant amounts of Ti and C. In this steel following irradiation at 80 C there were no visible precipitates, even though there are signicant amounts of C (60.12 wt.%, 60.56 at.%) and Ti (0.6 wt.%, 0.7 at.%), both of which are potent solution-strengthening elements. Assuming experimental hardening values provided for carbon by Irvine and coworkers [51] and for titanium by Gessel [52], we can estimate the matrixsoftening, although we are hampered by our earlier-mentioned lack of knowledge concerning the real carbon content of this rod. If we assume that the carbon is at the full specication value of 0.56 at.%, it can be shown in a simple FeCrNi steel that solutehardening would be a maximum of 141 kg/mm2 if carbon and titanium were fully in solution and linearly additive. At temperatures P450 C these two elements are being removed from solution, softening the matrix, but also contributing to positive hardening as precipitates. If we assume that we have some carbon and/or titanium remaining in solution, it must be recognized that superposition of solid solution and particle strengthening might not allow linear addition, but should be determined by a relationship such as expression (2) [53]. Given the uncertainty in carbon level and cumulative uncertainties in microstructural densities, the authors see no value to attempting to rene their calculations. It is sufcient to note that the difference in aP values for precipitates between our and earlier studies can easily be consistent with an interpretation involving solution strengthening followed by matrix softening. Additional attention should be paid to irradiation hardening observed at 5 dpa at 8090 C where Frank loops and black dot defects were observed to dominate. The calculation of irradiation hardening was rst made with coefcients aFL = 0.33 and aBD = 0.2, thought to be suitable for annealed steel. However, using these values the calculated value of Hl is about 27% lower than the experimentally determined 147 kg/mm2 (Fig. 10). It might be supposed that the cause of such discrepancy between prediction and experiment is connected with incorrect choices of barrier powers for these defects. Perhaps, however, additional factors may be involved. The selected value aBD = 0.2 for black dots is consistent with reported values for cluster defects in austenitic steels at low irradiation temperatures: 0.1460.2 [26,57]. However, it has been suggested that at relatively low irradiation temperatures, the power of defect barriers can change in comparison with that of clusters formed at higher irradiation temperatures, e.g., >300 C, especially for rigid barriers such as Frank loops. In Ref. [39] for 316SS irradiated by neutrons at 60100 C it was suggested that the average barrier strength aFL might be as high as 0.40.6. In Ref. [26] for 316LN steel irradiated at 90 C, the hardening coefcient for Frank loops was given as 0.5, signicantly larger than in the case of higher irradiation temperatures (a = 0.250.33 at >300 C for Cr18Ni9Ti, 316 and Cr18Ni10 steels [17,34,38,54]). For dislocation loops formed at lower temperatures an increase of barrier power might be attributed to formation of impurity atmospheres on the stacking faults of Frank loops. An increase of either irradiation or annealing temperature at 450550 C might signicantly weaken such atmospheres due to enhanced diffusion processes and/or formation of second phase precipitates. Therefore, it is not unreasonable to expect that the barrier power for Frank loops formed at 80 C is higher than 0.33. Note

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158

157

in Fig. 10, that if we assume for loops aFL = 0.5 similar to Refs. [26,39] and aBD = 0.2, then the estimation of Hl using Eqs. (1)(3) gives good agreement with the experimental result where Hlcalc = 141 kg/mm2. Hardening contributions from black dot defects serving as weak barriers can be estimated not only using a low aBD coefcient in the dispersed barrier scheme (1), but also using the FriedelKroupa Hirsch (FKH) hardening relationship for weak obstacles [55,58].

[6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16]

Dry 1=8M lbdN

2 =3

It is assumed that dispersed barrier model is more appropriate for barrier strengths of a = 0.251 [56], but the FKH model yields an adequate description when a < 0.25, i.e. about 1/4 of the impenetrable obstacle limit. However, in our case, calculation of hardening using the FKH model for black dots and using the dispersed barrier model for Frank loops with aFL = 0.5 was not successful. Such treatment yielded a hardening value 16% lower than experimentally observed. This result conrmed our decision to use only the barrier hardening model to estimate contributions from all defects. 4. Conclusions Studies of microstructural changes and mechanical properties for 12Cr18Ni9Ti steel were performed after irradiation for 20 years at a very low dpa rate in the WWR-K research reactor to doses of 0.001 and 5 dpa at temperature of 80 C and then isochronally annealed for 1 h in the range 4501050 C. The microstructure of radiation-induced defects signicantly changed as a result of heat treatment, leading to a decrease in radiation-induced hardening. It was found that a signicant amount of nely-dispersed secondary precipitates (TiC) was formed at dislocation loops after annealing at temperatures higher than 750 C, but these precipitates did not signicantly affect the materials strength. This behavior was explained by the fact that the elements Ti and C involved in precipitation are at the same time solution-strengthening agents. Removal of these elements from the matrix leads to softening of the matrix, compensating somewhat for enhanced precipitate hardening. The dispersed barrier hardening model with commonly accepted values of physical properties has been employed to successfully predict the hardening that develops across a wide variety of defect ensembles produced during irradiation at 80 C and during subsequent isochronal annealing. It also appears that the barrier power for Frank loops differs between the low temperature irradiation case and the post-irradiation heat treatment case. Acknowledgements This research was supported by ISTC Project K-172 and Project 0786/GF2. The authors would like to thank N.V. Sherbinina, V.M. Vereshak, and N.A. Brikotnina for their contribution to the research presented in this paper. References
[1] G.S. Was, Fundamentals of Radiation Materials Science, Springer, 2007. [2] P.J. Maziasz, C.J. McHargue, Int. Mater. Rev. 32 (1987) 90. [3] F.A. Garner, Void Swelling and Irradiation Creep in Light Water Reactor Environments, Understanding and Mitigating Ageing in Nuclear Power Plants; Materials and Operational Aspects of Plant Life Management (PLiM), Philip G. Tipping (Ed.), Woodhead Publishing Series in Energy: Number 4, 2010, pp. 308365 (Chapter 10). [4] A. Etienne, B. Radiquet, P. Pareige, J.-P. Massoud, C. Pokor, J. Nucl. Mater. 382 (2008) 6469. [5] S.T. Byrne, I. Wilson, R. Shogan, Microstructural characterization of bafe bolts, in: Proc. Fontevraud 5, Contribution of Materials Investigation to the

[17] [18] [19] [20]

[21] [22]

[23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36]

[37] [38] [39] [40] [41]

[42]

[43] [44] [45] [46] [47] [48]

Resolution of Problems Encountered in Pressurized Water Reactors, September 2327, 2002 (on CD format, no page numbers). F.A. Garner, M.B. Toloczko, J. Nucl. Mater. 251 (1997) 252261. P.J. Maziasz, J. Nucl. Mater. 205 (1993) 118145. S.J. Zinkle, P.J. Masiasz, R.E. Stoller, J. Nucl. Mater. 206 (1993) 266. T.R. Allen, J.I. Cole, C.L. Trybus, D.L. Porter, H. Tsai, F.A. Garner, E.A. Kenik, T. Yoshitake, J. Ohta, J. Nucl. Mater. 348 (2006) 148164. S.M. Bruemmer, E.P. Simonen, P.M. Scott, P.L. Andresen, G.S. Was, J.L. Nelson, J. Nucl. Mater. 274 (1999) 299314. D.J. Edwards, E.P. Simonen, S.M. Bruemmer, in: Microstructural Processes in Irradiated Materials 2000, vol. 650, MRS, Pittsburg, PA, 2001, Paper R2.7. D.J. Edwards, E.P. Simonen, S.M. Bruemmer, J. Nucl. Mater. 317 (2003) 1331. D.J. Edwards, E.P. Simonen, F.A. Garner, L.R. Greenwood, B.M. Oliver, S.M. Bruemmer, J. Nucl. Mater. 317 (2003) 3245. S.I. Porollo, Yu.V. Konobeev, A.M. Dvoriashin, A.N. Vorobjev, V.M. Krigan, F.A. Garner, J. Nucl. Mater. 307311 (2002) 339342. O.P. Maksimkin, K.V. Tsai, L.G. Turubarova, T. Doronina, F.A. Garner, J. Nucl. Mater. 329333 (2004) 625629. V.S. Neustroev, V.G. Dvoretzky, Z.E. Ostrovsky, V.K. Shamardin, G.A. Shimansky, Investigation of Microstructure and Mechanical Properties of 18Cr10NiTi Steel Irradiated in the Core of VVER-1000 Reactor, Effects of Radiation on Materials, in: 21st Intern. Symp., ASTM STP 1447, 2004, pp. 32 45. S.I. Porollo, A.M. Dvoriashin, Yu.V. Konobeev, A.A. Ivanov, S.V. Shulepin, F.A. Garner, J. Nucl. Mater. 359 (2006) 4149. N.I. Budylkin, T.M. Bulanova, E.G. Mironova, N.M. Mitrofanova, S.I. Porollo, V.M. Chernov, V.K. Shamardin, F.A. Garner, J. Nucl. Mater. 329333 (2004) 621624. O.P. Maksimkin, K.V. Tsai, L.G. Turubarova, T.A. Doronina, F.A. Garner, J. Nucl. Mater. 367370 (2007) 990994. F.A. Garner, N.I. Budylkin, Yu.V. Konobeev, S.I. Porollo, V.S. Neustroev, V.K. Shamardin, A.V. Kozlov, The Inuence of dpa Rate on Void Swelling of Russian Austenitic Stainless Steels, in: 11th International Conference on Environmental Degradation of Materials in Nuclear Power Systems Water Reactors, 2003, pp. 647656. T.M. Williams, ASTM STP 782 (1982) 166185. O.P. Maksimkin, L.G. Turubarova, N.V. Sherbinina, P.V. Chakrov, F.A. Garner, A.B. Johnson, Microstructural and Mechanical Studies of the Stainless Steel/ Aluminum Alloy Control Rod of the WWR-K Research Reactor, in: 10th International Conference on Environmental Degradation of Materials in Nuclear Power Systems Water Reactors, 2001 (issued on CD format, no page numbers). G.W. Groves, A. Kelly, Phil. Mag. 6 (1961) 15271529. M. Hiroki, M. Kiritani, J. Nucl. Mater. 239 (1996) 3441. M. Hiroki, S. Arai, Y. Satoh, M. Kiritani, J. Nucl. Mater. 255 (1998) 165173. N. Hashimoto, E. Wakai, J.P. Robertson, J. Nucl. Mater. 273 (1999) 95101. T. Sourmail, Mater. Sci. Technol. 17 (2001) 114. S.P. Vagin, B.D. Utkelbayev, P.V. Chakrov, J. Nucl. Mater. 233237 (1996) 1168 1171. V.Ph. Reutov, B.D. Utkelbayev, S.P. Vagin et al., Atomic Energy, 69, #3 (1990) 140142. W. Kesternich, R.V. Nandedkar, J. Nucl. Mater. 179181 (1991) 10151018. C.E. Danilov et al., Series Phys. Radiat. Mater. Sci. 3 (2002) 3135. J.T. Busby, M.M. Sowa, G.S. Was, E.P. Simonen, Phil. Mag. 85 (2005) 609617. Sh.Sh. Ibragimov, Radiation defects in metals and alloys. in: Radiation Defects in Metallic Crystals, Alma-Ata: Nauka Kaz. SSR, 1978. pp. 330. F.A. Garner, M.L. Hamilton, N.F. Panayotou, G.D. Johnson, J. Nucl. Mater. 103&104 (1981) 803. M. Ando, Y. Katoh, H. Tanigawa, A. Kohyama, T. Iwai, J. Nucl. Mater. 283287 (2000) 423427. J. Gan, D.J. Edwards, E.P. Simonen, S.M. Bruemmer, Microstructural evolution and hardening in 300-series stainless steels: comparison between neutron and proton irradiation, in: Proc. of 10th Intern. Symp. Environmental Degradation of Materials, NACE, Houston, TX, 2002. H.R. Higgy, F.H. Hammad, J. Nucl. Mater. 55 (1975) 177186. M. Grossbeck, P.J. Maziasz, A.F. Rowcliffe, J. Nucl. Mater. 191194 (1992) 808 812. N. Hashimoto, T.S. Byun, K. Farrell, J. Nucl. Mater. 295392 (2006) 351. J.I. Cole, S.M. Bruemmer, J. Nucl. Mater. 225 (1995) 5358. E.P. Simonen, D.J. Edwards, S.M. Bruemmer, Light-water reactor microstructural characterization from post-irradiation annealing behavior, in: Proc. Tenth Intern. Symp. on Environmental Degradation of Materials, NACE, Houston, TX, 2002. K. Asano, R. Katsura, M. Kodama, S. Nishimura, K. Fukuya, K. Nakata, PostIrradiation annealing effects on hardness and intergranular corrosion in Type 304 stainless steel, in: Proceedings of the 7th International Symposium on Environmental Degradation of Materials in Nuclear Power Systems-Water Reactors, NACE, International, 1995. J.T. Busby, G.S. Was, E.A. Kenik, J. Nucl. Mater. 302 (2002) 2040. C. Robertson, L. Boulanger, S. Poissonnet, J. Nucl. Mater. 271272 (1999) 102 105. G.E. Lucas, J. Nucl. Mater. 206 (1993) 287305. A.K. Seegar, in Second UN Conference on Peaceful Uses of Atomic Energy, vol. 6, United Nations, New York, 1958, p. 250. U.F. Kocks, Metall. Trans. 1 (1970) 11211143. J.T. Busby, M.C. Hash, G.S. Was, J. Nucl. Mater. 336 (2005) 267278.

158

K.V. Tsay et al. / Journal of Nuclear Materials 439 (2013) 148158 [53] E. Nembach, M. Martin, Acta Met. 28 (1980) 10691075. [54] V.S. Neustroev, Z.E. Ostrovsky, S.V. Belozerov, VANT; Series Phys. Radiat. Mater. Sci. 6 (2007) 7881. [55] J. Friedel, Dislocations, Pergamon, New York, 1964. [56] S.J. Zinkle, Y. Matsukawa, J. Nucl. Mater. 329333 (2004) 8896. [57] S.J. Zinkle, Radiat. Eff. Def. Solids. 148 (1999) 447477. [58] H. Fukushima, Y. Shimomura, J. Nucl. Mater. 205 (1992) 5967.

[49] M.N. Gusev, O.P. Maksimkin, O.V. Tivanova, N.S. Silnaygina, F.A. Garner, J. Nucl. Mater. 359 (2006) 258262. [50] V.S. Neustroev, E.V. Boev, F.A. Garner, J. Nucl. Mater. 367370 (2007) 935939. [51] K.J. Irvine, D.T. Llewellyn, F.B. Pickering, J. Iron Steel Institute 199 (1961) 153 175. [52] G.R. Gessel, Effects of Minor Alloying Additions on the Strength and Swelling Behavior of an Austenitic Stainless Steel, Oak Ridge National Laboratory Report ORNL/TM 6359, June 1978.

Potrebbero piacerti anche