Next Article in Journal
Study of the TIG Welding Process of Thin-Walled Components Made of 17-4 PH Steel in the Aspect of Weld Distortion Distribution
Next Article in Special Issue
Thermomechanical Properties of Neutron Irradiated Al3Hf-Al Thermal Neutron Absorber Materials
Previous Article in Journal
Comparative Filtration Performance of Composite Air Filter Materials Synthesized Using Different Impregnated Porous Media
Previous Article in Special Issue
Recent Progress in Gd-Containing Materials for Neutron Shielding Applications: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nanocluster Evolution in D9 Austenitic Steel under Neutron and Proton Irradiation

by
Suraj Venkateshwaran Mullurkara
1,2,3,
Akshara Bejawada
1,2,
Amrita Sen
1,4,
Cheng Sun
5,
Mukesh Bachhav
5 and
Janelle P. Wharry
1,*
1
School of Materials Engineering, Purdue University, West Lafayette, IN 47907, USA
2
Department of Metallurgical and Materials Engineering, Indian Institute of Technology Madras, Chennai 600036, India
3
Department of Mechanical Engineering & Materials Science, University of Pittsburgh, Pittsburgh, PA 15260, USA
4
Intel Corporation, Hillsboro, OR 97124, USA
5
Idaho National Laboratory, Idaho Falls, ID 83415, USA
*
Author to whom correspondence should be addressed.
Materials 2023, 16(13), 4852; https://doi.org/10.3390/ma16134852
Submission received: 3 May 2023 / Revised: 9 June 2023 / Accepted: 5 July 2023 / Published: 6 July 2023
(This article belongs to the Special Issue Advanced Characterization Techniques on Nuclear Fuels and Materials)

Abstract

:
Austenitic stainless steel D9 is a candidate for Generation IV nuclear reactor structural materials due to its enhanced irradiation tolerance and high-temperature creep strength compared to conventional 300-series stainless steels. But, like other austenitic steels, D9 is susceptible to irradiation-induced clustering of Ni and Si, the mechanism for which is not well understood. This study utilizes atom probe tomography (APT) to characterize the chemistry and morphology of Ni–Si nanoclusters in D9 following neutron or proton irradiation to doses ranging from 5–9 displacements per atom (dpa) and temperatures ranging from 430–683 °C. Nanoclusters form only after neutron irradiation and exhibit classical coarsening with increasing dose and temperature. The nanoclusters have Ni3Si stoichiometry in a Ni core–Si shell structure. This core–shell structure provides insight into a potentially unique nucleation and growth mechanism—nanocluster cores may nucleate through local, spinodal-like compositional fluctuations in Ni, with subsequent growth driven by rapid Si diffusion. This study underscores how APT can shed light on an unusual irradiation-induced nanocluster nucleation mechanism active in the ubiquitous class of austenitic stainless steels.

1. Introduction

Austenitic stainless steels (SS) such as 304 SS and 316 SS are the most ubiquitous structural alloys in light water reactor (LWR) nuclear power plants [1,2]. But these alloys are susceptible to irradiation hardening and embrittlement due to the formation of voids, dislocation loops, and stacking fault tetrahedra [3,4,5,6,7,8,9,10]. These alloys also exhibit poor creep resistance, especially at elevated temperatures as would be required in advanced Generation IV nuclear reactors such as supercritical water-cooled reactors (SCWR) and sodium-cooled fast reactors (SFR) [11,12,13,14]. Consequently, 316 SS has been modified with a higher Ni/Cr ratio and higher Si and Ti concentrations for enhanced radiation tolerance [15,16,17,18,19] and high-temperature creep strength [20]. This modified material, known as alloy D9, has since been included in several neutron irradiation campaigns at the Fast Flux Test Facility (FFTF) [17,18,19,21], High Flux Isotope Reactor (HFIR) [15,22], Experimental Breeder Reactor-II (EBR-II) [16,23,24], and Advanced Test Reactor (ATR) [25].
The FFTF experiments focused on the irradiation-induced mechanical property evolution of D9. Over irradiation temperatures of 390–457 °C, fracture toughness degrades with increasing dose from ~45 dpa to ~135 dpa (corresponding to neutron fluences 9–27 × 1022 n/cm2) [18]. More generally, though, irradiation temperatures over the range of 395–686 °C have relatively little influence on post-irradiation mechanical properties, including reduction in ductility [21]. Meanwhile, the irradiation experiments in HFIR and EBR-II focus more on the irradiated microstructure evolution of D9. Maziasz [15,22] specifically studied void swelling of various process heats of D9 and found that swelling is suppressed at irradiation temperatures below 500 °C, even at doses as high as 34–57 dpa. Void incubation appears to give way to steady-state swelling when G-phase precipitation begins [16]. This kinetic relationship between voids and G-phases is due to the role of P and Si [22] in stabilizing fine, high-density precipitates, thus extending the void incubation time [15,23]. The sizes of both voids and G-phase precipitates increase with increasing irradiation temperature from 427 to 593 °C [23].
Despite these aforementioned irradiation effects studies on alloy D9, the irradiated microstructure and mechanical properties had not been directly related until Chen and colleagues [26] conducted nanoindentation and transmission electron microscopy (TEM) characterization of several ATR-irradiated D9 specimens. They observed conventional irradiation hardening behaviors as a function of dose. The hardening is directly related to the Frank loops, cavities, and irregular Ni–Si nanoclusters through the Orowan dispersed barrier hardening model (although cavities are ultrafine and have a relatively trivial contribution to overall hardening). However, the highest degree of hardening, which occurs at the highest irradiation dose, also occurs at the highest irradiation temperature. This is somewhat unusual, since the extent of hardening typically decreases with temperature. Due to the limited number of specimens, the role of temperature and dose could not be deconvoluted. However, the authors theorize that the Ni–Si nanoclusters may be able to explain the hardening behaviors.
A unique aspect of the irradiated microstructure that remains to be fully explored is this Ni–Si nanoclustering. Although Chen et al. [26] characterized these nanoclusters, their study notes specific gaps in the understanding of the chemistry and morphologies of these nanoclusters, which they believe will influence the mechanical behaviors and irradiation hardening. Chen and colleagues [26] note that they characterized the nanoclusters using TEM with energy dispersive X-ray spectroscopy (EDS), but, given the ~few nm resolution of TEM, and the through-thickness nature of EDS composition measurements, greater accuracy of the Ni–Si nanocluster compositions is warranted. Hence, the use of atom probe tomography (APT) will complement Chen’s work by providing three-dimensional, atomic resolution reconstruction of irradiation-induced nanoclusters in alloy D9.
The present study accesses the identical neutron-irradiated specimens of D9 previously characterized by Chen et al. [26]. The as-received archival material from the original D9 ingot is studied as a control specimen, and a piece of this ingot is also irradiated with protons to a dose and temperature that fall within the range of doses and temperatures of the neutron irradiations studied. APT characterization of all specimens reveals a fine dispersion of Ni–Si nanoclusters in the neutron-irradiated specimens. The role of irradiation dose and temperature on nanocluster morphology is discussed. The use of APT provides unique insight into the structure of nanoclusters, which can shed light on the mechanisms of Ni–Si nanocluster nucleation in irradiated austenitic steels.

2. Materials and Methods

2.1. Alloy Preparation and Irradiation

An ingot of Alloy D9 was obtained from Oak Ridge National Laboratory (Oak Ridge, TN, USA). The alloy composition is provided in Table 1 using APT and compared to scanning electron microscopy (SEM) EDS measurements from ref. [26] and the nominal composition [25]. While all composition measurements are in generally good agreement, especially for the major alloying species, APT is sensitive to minor elements with low concentrations, but is more susceptible to local heterogeneities. The D9 specimens were machined into discs of 3 mm diameter and ~500 µm thickness using electrical discharge machining (EDM). The disc faces were polished using SiC paper through P4000 grit to a thickness of ~250 µm.
Discs were loaded into capsules for neutron irradiation in the Advanced Test Reactor (ATR) at Idaho National Laboratory (INL) (Idaho Falls, ID, USA). The capsules were irradiated in the ATR East Flux Trap, which has a maximum fast flux of ~9.7 × 1013 n/cm2-s (E > 1 MeV), during cycles 143A through 150B; the specimen loading configuration is described in ref. [25]. Three D9 neutron-irradiated specimens were selected for this study, as shown in Table 2. The specimen geometries were 3 mm discs of ~200 µm thickness, so the neutrons can be assumed to deliver a homogeneous fluence throughout the specimen volume. These specimens accumulated total neutron fluences of 1.65 × 1022, 2.69 × 1022, and 3.07 × 1022 neutrons/cm2, corresponding to damage doses of 5, 8, and 9 displacements per atom (dpa), respectively. The irradiation temperatures were determined to be 448, 430, and 683 °C, respectively [27], based on ABAQUS calculations with input from reactor instrumentation and SiC melt wires, following methods described in ref. [28] with temperature uncertainties of approximately ±20 °C.
A proton irradiation was also conducted to conditions falling within the dose and temperature range bounded by the neutron irradiations, as shown in Table 2. For proton irradiation, a specimen from the same ingot of alloy D9 was cut into a 2 mm × 2 mm × 20 mm matchstick using EDM. The surface to be irradiated was mechanically polished through P4000 grit SiC paper. Subsequently, the specimen was electropolished in a 10% perchloric acid + 90% methanol solution maintained at −33 ± 3 °C with a Pt mesh electrode. The specimen was included in an irradiation stage with several other specimens of various steels, the irradiation results from which have been reported in refs. [29,30,31,32]. The proton–target interaction was calculated using SRIM 2013 in Quick Kinchin–Pease mode, and the dose of 7 dpa was determined at half the depth to the damage peak, or ~10 µm below the irradiated surface, as described in ref. [30]. The irradiation temperature was calibrated using thermocouples spot-welded to the specimen surfaces, and the temperature was tracked throughout the irradiation using a non-contact 2D infrared thermal pyrometer. The reader is referred to ref. [29] for comprehensive details about the proton irradiation experiment setup, control, and monitoring.

2.2. Atom Probe Tomography

APT needles were fabricated using established focused ion beam (FIB) milling procedures on a Thermo Fisher Scientific Quanta 3D FEG dual-beam SEM/FIB (Waltham, MA, USA) at the Center for Advanced Energy Studies (CAES) [33]. APT needle fabrication targeted grain interiors, aiming to avoid grain boundaries or near-boundary regions with potential denuded zones. A ~200 nm Pt protective layer was deposited onto the identified regions of interest, then trench cuts were made adjacent to the Pt layer. Using conventional under-cutting and extraction procedures, a lamella was lifted out and welded onto Si posts on an APT microtip array coupon. The microtips were subsequently milled using an annular pattern with sequentially decreasing FIB beam currents. Final beam energies of 5 keV and 2 keV were used to sharpen each needle to a diameter of <100 nm, while minimizing FIB damage on needle surfaces. In the proton-irradiated specimens, a larger wedge lift-out was extracted, such that the sharpened tips of the needles would be situated ~1 µm below the original irradiated surface, which would correspond to an actual irradiation dose of ~6 dpa.
The needles were tested in a CAMECA (Madison, WI, USA) Local Electrode Atom Probe (LEAP) 5000X HR, also at CAES, operating in laser pulse mode. The LEAP run parameters for each needle are summarized in Table 3; across all needles, the laser energy ranged from 0.06–0.14 NJ, pulse rate ranged from 200–250 kHz, specimen base temperature ranged from 39.5–60.5 K, and detection rate ranged from 0.5–1%. APT data was reconstructed using the proprietary CAMECA Integrated Visualization and Analysis Software (IVAS) version 3.8.4. A minimum of two needles were run from each specimen condition, and a minimum of ~1,000,000 ions were collected from each needle in order for the data set to be considered statistically sound. During reconstruction, the ranged Ni ions were determined using the peaks at both 29 and 30 Da in the time-of-flight mass spectrum. Although the 29 Da peak overlaps with Fe, this peak accounts for >68% of all Ni ions and only 0.3% of all Fe ions, based on natural isotopic abundances [34]. Thus, despite the higher bulk concentration of Fe than Ni in the D9 alloy, only a relatively negligible amount of Fe ions would be ranged as Ni ions in the 29 Da peak.
Cluster analysis was conducted within IVAS using the maximum separation method [35] for Si ions. The maximum allowable distance between solute atoms in a cluster, dmax, and the minimum number of ions that must satisfy the dmax condition to be considered a cluster, Nmin, were varied for each needle following recommendations from refs. [36,37,38,39]. The values of dmax and Nmin were consistent within each irradiation condition and varied slightly across irradiation conditions, as shown in Table 3. Cluster analysis provided compositional and dimensional information for each cluster, and the total number of clusters identified per needle. The average nanocluster size was determined from the Guinier diameter, DG, as [40]:
D G = 2 5 3 R g
where Rg is the overall radius of gyration, which is estimated from the gyration radii in the x, y, and z directions for each cluster [41]:
R g = R g x 2 + R g y 2 + R g z 2
The nanocluster number density was determined by:
N n c = N C V T
where VT is the analyzed volume of a given needle, and NC is the number of clusters identified using the maximum separation method.

3. Results

3.1. As-Received and Proton-Irradiated

Three APT needles were reconstructed from the as-received material, and five from the proton-irradiated specimen. All of the reconstructions show homogeneous distributions of all major alloying elements, Fe, Cr, Mn, Ni, and Si, in both the as-received and proton-irradiated specimens, as shown in Figure 1. This degree of homogeneity is expected in the as-received specimen, since previous characterization studies of this heat of D9 [26] show no evidence of pre-existing nanoclusters. The Ti distribution is also homogeneous in the as-received specimen but appears concentrated toward a side of the proton-irradiated specimen. This skewing is likely an artifact of the Ti peak in the APT mass spectrum (Da~45) being significantly larger than that of the other elements studied; APT laser settings optimized for elements such as Fe, Cr, and Ni, which have peaks in the range Da ~20–30, are not as ideal for the larger Da value of Ti.
The proton-irradiated needles reveal no evidence of irradiation-induced nanoclustering or nanoprecipitation. It is also worth noting that no chemical segregation to features such as grain boundaries, dislocation lines, or dislocation loops is observed in any of the as-received or proton-irradiated needles. However, the APT needle preparation process did not specifically aim to capture grain boundaries. Additionally, the likelihood of capturing dislocation loops or lines in the nanoscopic volume of an APT needle is low, and such features can also pop out to the surface of needles upon application of the APT laser pulse. Hence, no conclusive statements about nano-segregation behaviors can be drawn.

3.2. Neutron-Irradiated

A total of six, two, and five APT needles were reconstructed from the 5, 8, and 9 dpa neutron-irradiated specimens, respectively, with representative reconstructions shown in Figure 2. At all neutron irradiation conditions, a fine dispersion of Ni–Si rich nanoclusters forms throughout the material. The other alloying elements of Fe, Cr, Mn, and Ti are not enriched in the nanoclusters and are homogeneously distributed, although the Ti distribution appears skewed for the 8 dpa condition, likely due to the same mass spectrum artifact noted in the proton-irradiated specimen. The average Ni concentration in nanoclusters increases with irradiation dose, from 22.53 to 32.74 at%, as shown in Table 4. The Si concentration in nanoclusters also increases with irradiation dose, maintaining a statistically invariant Ni/Si ratio in nanoclusters ranging from 3.08–3.25.
The average nanocluster morphologies are also summarized in Table 4 and presented in Figure 3a. The average nanocluster size increases from 5.3 ± 3.8 nm at 5 dpa, through to 8.7 ± 8.8 nm at 9 dpa. Meanwhile, the nanocluster number density exhibits a statistically significant decrease with increasing dose, from 2.45 ± 0.99 × 1024 m−3 at 5 dpa, through to 1.31 ± 0.31 × 1024 m−3. This coarsening behavior is reflected by an increase in nanocluster volume fraction, which is 0.19, 0.21, and 0.45, at 5, 8, and 9 dpa, respectively. The volume fraction increase is only statistically significant between the 8 dpa, 430 °C and the 9 dpa, 683 °C conditions, suggesting that volume fraction increase may be most strongly influenced by temperature. The nanocluster size distribution, as shown in Figure 3b, is positively skewed at all doses. However, the magnitude of skewness decreases with increasing dose, which is also consistent with nanocluster coarsening.
Nanoclusters appear to have a core–shell type structure in which the Ni-rich core is surrounded by a Si-rich shell, as shown in Figure 4. While this core–shell structure seems to appear at all irradiation conditions (Figure 4a,b), it is most clearly observed in the 9 dpa, 683 °C condition, as shown in Figure 4c. Isoconcentration surfaces are defined for concentrations near the average nanocluster concentration of 10 at% Si and 30 at% Ni (i.e., indicative of the Ni3Si stoichiometry). At the lower irradiation temperatures, Ni isoconcentration surfaces cannot be defined, confirming that the nanoclusters do not achieve Ni3Si stoichiometry. At the higher irradiation temperature, the isoconcentration surfaces for Ni are smaller than for Si, as shown in Figure 4c, which is further indicative of the core–shell structure. The nanoclusters are more complex in shape and morphology at the 9 dpa, 683 °C condition than at the lower doses and temperatures; the isosurfaces appear nearly interconnected like spinodal-type phases.

3.3. Radiation Induced Segregation

One APT needle from the 8 dpa, 430 °C neutron-irradiated specimen passed over a grain boundary. The RIS behavior across this boundary is summarized in Figure 5. The boundary is depleted of Fe, Cr, Mo, and Mn; limited depletion of Ti also occurs. The only two species that enrich at the grain boundary are Ni and Si, consistent with the species that form into irradiation-induced nanoclusters. The Ni/Si ratio at the grain boundary is ~4.3, but little can be inferred from this composition ratio, since the as-received grain boundary chemistry has not been captured in any of the APT needles. It is also worth noting that grain boundary RIS tends to be sensitive to grain boundary misorientation angle [42,43], and that grain boundary segregation patterns are also sensitive to the orientation and associated stress fields surrounding dislocations that comprise grain boundaries [44].

4. Discussion

Irradiation-induced Ni–Si nanoclusters or γ’-phase nanoprecipitates have long been known to form in austenitic stainless steels [4,7,26,45,46,47,48,49,50,51,52,53,54]. Zinkle and colleagues’ 1993 compilation [55] of irradiation-induced precipitation behaviors in legacy specimens of solution-annealed 316 SS suggests that the γ’ precipitate phase forms within a limited range of temperatures, from 400–500 °C, and doses exceeding ~40 dpa. They also note that fine γ’ precipitates can form at higher temperatures after only a few dpa, but subsequently dissolve at higher doses. Table 5 summarizes a more comprehensive set of observations of irradiation-induced Ni–Si nanoclusters from the archival literature in a wider range of commercial and model austenitic stainless steels; the cluster sizes, number densities, compositions, and irradiation conditions are tabulated, if reported. Most notably, this summary of Ni–Si nanoclusters suggests that Ni–Si nanoclusters and γ’ precipitates can exist over a considerably wider irradiation temperature and dose envelope than previously suggested. That is, there is evidence of Ni–Si nanoclustering in 304 SS, 316 SS, D9, and a model austenitic steel at temperatures ranging from 200–683 °C and doses ranging from 2–80 dpa.
Although the summarized data are limited and were collected over nonsystematic irradiation conditions and from different process heats of alloys, a few general trends can be extracted. First, the size and number density evolution of nanoclusters may suggest coarsening with increasing dose and/or temperature; this can be seen visually with bubble plots, as shown in Figure 6a,b. Quantitative measurements of nanocluster sizes and densities in the present study are self-consistent and indicative of coarsening with dose and/or temperature. However, by comparison to other studies that use TEM, the use of APT herein resolves smaller nanoclusters at a higher number density. It is also worth noting that nanocluster number densities in 304 SS tend to be several orders of magnitude greater than those in 316 SS, although it is unclear whether this is due to actual differences in bulk composition or solute chemistry between the alloys, or whether this is merely an artifact of the limited data set available. The nanocluster sizes and densities appear insensitive to bulk Si concentration, although the summarized studies do not use consistent methods of measuring Si concentration.
At elevated doses, Ni–Si nanoclusters are generally considered to be γ’-phase precipitates, which have simple cubic structures with Ni3Si stoichiometry [54,56] and lattice parameters of 0.3505 nm [48]. The current study is consistent with nearly all other reports in archival literature, showing that nanoclusters have approximately Ni3Si stoichiometry (i.e., Ni/Si ratio~3), as shown in Table 5. The stoichiometry appears insensitive to irradiation dose and temperature, as shown in Figure 6c, and to the measurement technique (APT or TEM). One notable case in which the nanoclusters deviate from Ni3Si stoichiometry is the Fe-17Cr-12Ni-1Si model steel irradiated with 6.4 MeV Fe3+ ions at 200 °C, from Fukumoto et al. [53]. They characterize the nanoclusters at three depths ranging from 500 nm (corresponding to 2.0 dpa) through the damage peak at 1500 nm (corresponding to 6.7 dpa); over this span, the nanocluster Ni/Si ratio ranges from 2.1–2.5. This is the only study summarized in Table 5 that focuses on a high-purity model alloy and the only study in which ions induce Ni–Si nanocluster nucleation; these unique aspects may contribute to the deviation from the Ni3Si nanocluster stoichiometry.
The Ni–Si precipitation mechanism remains unclear. Historically, Ni–Si precipitates have been thought to initially nucleate at interstitial sinks [49,57], suggesting an interstitial diffusion-driven nucleation mechanism. With increasing fluence, Ni–Si precipitates nucleate in the matrix, due to RIS eventually causing compositional segregations at all defect sinks [55,58,59,60]. But RIS of Ni in austenitic stainless steels is ascribed to diffusion via the vacancy flux, rather than via the interstitial flux [61,62]. When combined, these theories might suggest that the transport of Ni to nanoclusters is at first interstitial-mediated but evolves to be vacancy-mediated. While this is plausible, it requires significant co-evolution of sink strengths of multiple microstructural features in the system. The possible transition from interstitial- to vacancy-mediated clustering could also be consistent with rate theory studies in pure Ni–Si that show that irradiation-induced γ’ growth has a lower activation enthalpy at irradiation temperatures > 590 °C than at temperatures ranging from 475–570 °C [63].
Alternatively, the Ni core–Si shell structure of the nanoclusters observed in the present work may be able to shed light on possible nucleation mechanisms. This core–shell structure has not been reported in other studies; although Toyama et al. [7] describe a Mn core–P shell type structure of Ni–Si rich precipitates in 304 SS fuel wrapper plates (24 dpa, 300 °C), Ni and Si do not themselves form a core–shell structure in that work. The Ni–Si core–shell structure is consistent with the evolution of nanoclusters during post-irradiation annealing as reported in refs. [47,53,54]. For example, Jiao et al. [47] conduct post-irradiation annealing of specimens from a 304L SS control rod (5.9 dpa, maximum temperature 288 °C during full insertion). While the as-irradiated Ni–Si nanoclusters do not exhibit a core–shell structure, Si diffuses more rapidly than Ni out of the clusters upon post-irradiation annealing [47], leaving a Ni core-like structure after partial annealing. Similarly, Fukumoto et al. [53] observe that the Ni/Si ratio of ion irradiation-induced nanoclusters increases with post-irradiation annealing temperature, suggesting that Si diffuses out of nanoclusters more rapidly than does Ni. Even grain boundary segregation of Ni and Si exhibit similar post-irradiation annealing phenomena, in which Si recovers more readily than Ni [54]. It is possible that similar phenomena may be occurring in the present study, in which Si may be diffusing out of nanocluster peripheries after irradiation, artificially creating the appearance of the Si core–Ni shell structure. It may also be plausible that these nanoclusters form this core–shell structure as a result of segregation to dislocation loops, as their number density is inconsistent with Chen and colleagues’ measurements of dislocation loop densities in the identical material [26].
In austenitic stainless steel or γ-Fe, Si [64] has a higher diffusion coefficient than Ni [65,66]. Additionally, the activation energy for Ni diffusion (via vacancies or interstitials) [67] is higher than that for Si diffusion [68]. Thus, since Ni is a slower diffuser than Si, a Ni core with Si shell structure suggests that nanocluster cores do not initially nucleate by diffusion. Rather, the Ni cores may form through local compositional fluctuations [69], which can be likened to spinodal decomposition; this can also explain the interconnected and spinodal-like appearance of the isosurfaces (Figure 4b,c). A similar nucleation mechanism has been reported in U-50 wt% Zr under irradiation, in which a spinodal-like decomposition evolves into binodal precipitate growth [70,71]. Consequently, a modified U–Zr phase diagram has been proposed with extended spinodal and binodal curves; similarly, the Fe–Ni phase diagram has spinodal and binodal curves at the FeNi3 stoichiometry, which makes the hypothesized mechanism plausible. The nanoclusters are known to be coherent with the matrix, and their growth is believed to be promoted by coherency strain [53]. This, too, is consistent with a possible spinodal-like nucleation. The coherency leads to a high sink strength at nanocluster-matrix interfaces, and subsequent rapid radiation-enhanced diffusion of Si will result in further growth of the nanoclusters. The key role of Ni in driving the Ni–Si precipitation is similar to the dominant role of Ni in initiating G-phase precipitation in reactor pressure vessel (RPV) steels [72].
Finally, it is also worth commenting on the absence of Ni–Si nanoclusters in the proton-irradiated specimen herein. This observation is consistent with previous studies showing that ion [73,74] and proton [69,75,76,77] irradiation cannot replicate neutron irradiation-induced nanoclustering of a’ phases (sometimes described as Cr-rich nanoclusters) and Ni–Mn–Si–P (i.e., G-phase) nanoclusters in bcc Fe–Cr alloys, Fe–Cr–Al alloys [78], and oxide dispersion strengthened (ODS) steels. The exact mechanisms for this behavior are unknown but have been theorized to be associated with the accelerated irradiation damage rate of protons and ions compared to neutrons. However, ions may be able to induce Ni–Si nanoclustering in austenitic steels, considering the observations of nanoclusters due to 6.4 MeV Fe3+ ion irradiation in a study from Fukumoto et al. [53]. Nevertheless, the implanted ions may alter precipitation behavior [79].

5. Conclusions

APT was used to investigate austenitic stainless steel D9 irradiated with neutrons to 5 dpa at 448 °C, 8 dpa at 430 °C, and 9 dpa at 683 °C, and with 2 MeV protons to 7.0 dpa at 500 °C. Irradiation-induced nucleation of Ni–Si nanoclusters are consistently observed in all neutron-irradiated specimens, but they do not nucleate under proton irradiation. The nanoclusters coarsen with increasing neutron irradiation dose and temperature, but the Ni/Si ratio remains relatively constant near Ni3Si (i.e., γ’-phase) stoichiometry. The stability of these nanoclusters at the highest irradiation temperature, 682.5 °C, suggests that Ni–Si nanoclusters may be able to exist over a wider irradiation temperature range than previously thought. The observed nanoclusters exhibit a core–shell structure with a Ni-rich core and a more diffuse Si shell. Although a core–shell structure has not previously been reported for Ni–Si nanoclusters, it is consistent with post-irradiation annealing in the archival literature, which shows Si recovers more readily than Ni from nanoclusters. Because Ni is a slower diffuser than Si, the core–shell structure suggests that the Ni cores may nucleate through local compositional fluctuations akin to a spinodal-like phase instability, then subsequently grow through rapid Si diffusion to the core–matrix interface.

Author Contributions

Conceptualization, M.B., C.S. and J.P.W.; methodology, M.B., A.S. and J.P.W.; validation, A.S. and M.B.; formal analysis, S.V.M. and A.B.; investigation, A.S. and M.B.; resources, M.B., C.S. and J.P.W.; data curation, S.V.M., A.B. and A.S.; writing—original draft preparation, J.P.W.; writing—review and editing, S.V.M., A.B., A.S., C.S. and M.B.; visualization, S.V.M., A.B. and J.P.W.; supervision, M.B. and J.P.W.; project administration, J.P.W.; funding acquisition, M.B., C.S. and J.P.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the US Department of Energy, Office of Nuclear Energy, through Nuclear Science User Facilities (NSUF) rapid turnaround experiment 17-1081. J.P.W., A.S., S.V.M., and A.B. were supported by the National Science Foundation award DMR-1752636. A.S. was partially supported by the US Nuclear Regulatory Commission grant 31310021M0035. S.V.M. and A.B. were partially supported by the Purdue-India Partnership Undergraduate Research Experience (PURE).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is available upon request to the corresponding authors.

Acknowledgments

The authors thank Tiayni Chen of Oregon State University for the rich technical discussions; Yaqiao Wu, Ching-Heng Shiau, Yu Lu, Jatuporn Burns, Megha Dubey, and the staff at CAES for their support with specimen handling, FIB, and APT; and Ovidiu Toader, Fabian Naab, and Zhijie (George) Jiao at the Michigan Ion Beam Laboratory for their assistance with proton irradiation setup and monitoring.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Zinkle, S.J.; Was, G.S. Materials challenges in nuclear energy. Acta Mater. 2013, 61, 735–758. [Google Scholar] [CrossRef]
  2. Zinkle, S.J.; Busby, J.T. Structural materials for fission and fusion energy. Mater. Today 2009, 12, 12–19. [Google Scholar] [CrossRef]
  3. Huang, Y.; Wiezorek, J.M.K.; Garner, F.A.; Freyer, P.D.; Okita, T.; Sagisaka, M.; Isobe, Y.; Allen, T.R. Microstructural characterization and density change of 304 stainless steel reflector blocks after long-term irradiation in EBR-II. J. Nucl. Mater. 2015, 465, 516–530. [Google Scholar] [CrossRef]
  4. Van Renterghem, W.; Al Mazouzi, A.; Van Dyck, S. Influence of post irradiation annealing on the mechanical properties and defect structure of AISI 304 steel. J. Nucl. Mater. 2011, 413, 95–102. [Google Scholar] [CrossRef]
  5. Pokor, C.; Brechet, Y.; Dubuisson, P.; Massoud, J.-P.; Averty, X. Irradiation damage in 304 and 316 stainless steels: Experimental investigation and modeling. Part II: Irradiation induced hardening. J. Nucl. Mater. 2004, 326, 30–37. [Google Scholar] [CrossRef]
  6. Cole, J.I.; Bruemmer, S.M. Post-irradiation deformation characteristics of heavy-ion irradiated 304L SS. J. Nucl. Mater. 1995, 225, 53–58. [Google Scholar] [CrossRef] [Green Version]
  7. Toyama, T.; Nozawa, Y.; Van Renterghem, W.; Matsukawa, Y.; Hatakeyama, M.; Nagai, Y.; Al Mazouzi, A.; Van Dyck, S. Irradiation-induced precipitates in a neutron irradiated 304 stainless steel studied by three-dimensional atom probe. J. Nucl. Mater. 2011, 418, 62–68. [Google Scholar] [CrossRef]
  8. Wiezorek, J.M.K.; Huang, Y.; Garner, F.A.; Freyer, P.D.; Sagisaka, M.; Isobe, Y.; Okita, T. Transmission electron microscopy of 304-type stainless steel after exposure to neutron flux and irradiation temperature gradients. Microsc. Microanal. 2014, 20, 1822–1823. [Google Scholar] [CrossRef] [Green Version]
  9. Mao, K.S.; French, A.J.; Liu, X.; Wu, Y.; Giannuzzi, L.A.; Sun, C.; Dubey, M.; Freyer, P.D.; Tatman, J.K.; Garner, F.A.; et al. Microstructure and microchemistry of laser welds of irradiated austenitic steels. Mater. Des. 2021, 206, 109764. [Google Scholar] [CrossRef]
  10. Mao, K.S.; Sun, C.; Shiau, C.-H.; Yano, K.H.; Freyer, P.D.; El-Azab, A.A.; Garner, F.A.; French, A.; Shao, L.; Wharry, J.P. Role of cavities on deformation-induced martensitic transformation pathways in a laser-welded, neutron irradiated austenitic stainless steel. Scr. Mater. 2020, 178, 1–6. [Google Scholar] [CrossRef]
  11. Jayakumar, T.; Mathew, M.D.; Laha, K.; Sandhya, R. Materials development for fast reactor applications. Nucl. Eng. Des. 2013, 265, 1175–1180. [Google Scholar] [CrossRef]
  12. Mansur, L.K.; Rowcliffe, A.F.; Nanstad, R.K.; Zinkle, S.J.; Corwin, W.R.; Stoller, R.E. Materials needs for fusion, Generation IV fission reactors and spallation neutron sources—Similarities and differences. J. Nucl. Mater. 2004, 329–333, 166–172. [Google Scholar] [CrossRef]
  13. Crawford, D.C.; Porter, D.L.; Hayes, S.L. Fuels for sodium-cooled fast reactors: US perspective. J. Nucl. Mater. 2007, 371, 202–231. [Google Scholar] [CrossRef]
  14. Karthik, V.; Murugan, S.; Parameswaran, P.; Venkiteswaran, C.N.; Gopal, K.A.; Muralidharan, N.G.; Saroja, S.; Kasiviswanathan, K.V. Austenitic Stainless Steels for Fast Reactors -Irradiation Experiments. Prop. Eval. Microstruct. Stud. Energy Procedia 2011, 7, 257–263. [Google Scholar] [CrossRef] [Green Version]
  15. Maziasz, P.J. Overview of microstructural evolution in neutron-irradiated austenitic stainless steels. J. Nucl. Mater. 1993, 205, 118–145. [Google Scholar] [CrossRef]
  16. Maziasz, P.J. Formation and stability of radiation-induced phases in neutron-irradiated austenitic and ferritic steels. J. Nucl. Mater. 1989, 169, 95–115. [Google Scholar] [CrossRef]
  17. Garner, F.A.; Toloczko, M.B.; Sencer, B.H. Comparison of swelling and irradiation creep behavior of fcc-austenitic and bcc-ferritic/martensitic alloys at high neutron exposure. J. Nucl. Mater. 2000, 276, 123–142. [Google Scholar] [CrossRef]
  18. Huang, F.H. Comparison of fracture behavior for low-swelling ferritic and austenitic alloys irradiated in the Fast Flux Test Facility (FFTF) to 180 DPA. Eng. Fract. Mech. 1992, 43, 733–748. [Google Scholar] [CrossRef] [Green Version]
  19. Pitner, A.L.; Gneiting, B.C.; Bard, F.E. Irradiation Performance of Fast Flux Test Facility Drivers Using D9 Alloy. Nucl. Technol. 1995, 112, 194–203. [Google Scholar] [CrossRef]
  20. Latha, S.; Mathew, M.D.; Parameswaran, P.; Rao, K.B.S.; Mannan, S.L. Thermal creep properties of alloy D9 stainless steel and 316 stainless steel fuel clad tubes. Int. J. Press. Vessel. Pip. 2008, 85, 866–870. [Google Scholar] [CrossRef]
  21. Cannon, N.; Huang, F.; Hamilton, M. Transient and Static Mechanical Properties of D9 Fuel Pin Cladding and Duct Material Irradiated to High Fluence, in: Effects of Radiation on Materials. In Proceedings of the 15th Symposium on Effects of Radiation on Materials, Nashville, TN, USA, 17–21 June 1990; pp. 1071–1082. [Google Scholar] [CrossRef]
  22. Maziasz, P.J. Void swelling resistance of phosphorus-modified austenitic stainless steels during HFIR irradiation at 300–500C to 57 dpa. J. Nucl. Mater. 1993, 200, 90–107. [Google Scholar] [CrossRef]
  23. Lee, E.H.; Mansur, L.K. Relationships between phase stability and void swelling in Fe-Cr-Ni alloys during irradiation. Metall. Trans. A 1992, 23, 1977–1986. [Google Scholar] [CrossRef]
  24. Mansur, L.K.; Grossbeck, M.L. Mechanical property changes induced in structural alloys by neutron irradiations with different helium to displacement ratios. J. Nucl. Mater. 1988, 155–157, 130–147. [Google Scholar] [CrossRef]
  25. MacLean, H.J.; Sridharan, K.; Hyde, T.A. Irradiation Test Plan for the ATR National Scientific User Facility; University of Wisconsin Pilot Project; Idaho National Laboratory: Idaho Falls, ID, USA, 2008. [Google Scholar] [CrossRef] [Green Version]
  26. Chen, T.; He, L.; Cullison, M.H.; Hay, C.; Burns, J.; Wu, Y.; Tan, L. The correlation between microstructure and nanoindentation property of neutron-irradiated austenitic alloy D9. Acta Mater. 2020, 195, 433–445. [Google Scholar] [CrossRef]
  27. Wilson, S. As-Run Thermal Analysis for the University of Wisconsin Experiment (ECAR-3186); Idaho National Laboratory: Idaho Falls, ID, USA, 2012.
  28. Guillen, D.P.; Wharry, J.P.; Housley, G.; Hale, C.D.; Brookman, J.; Gandy, D.W. Irradiation Experiment Design for the Evaluation of PM-HIP Alloys for Nuclear Reactors. Nucl. Eng. Des. 2023, 402, 112114. [Google Scholar] [CrossRef]
  29. Penisten, J. The Mechanism of Radiation-Induced Segregation in Ferritic-Martensitic Steels. Ph.D. Dissertation, University of Michigan, Ann Arbor, MI, USA, 2012. [Google Scholar]
  30. Wharry, J.P.; Was, G.S. A systematic study of radiation-induced segregation in ferritic-martensitic alloys. J. Nucl. Mater. 2013, 442, 7–16. [Google Scholar] [CrossRef] [Green Version]
  31. Wharry, J.P.; Jiao, Z.; Was, G.S. Application of the inverse Kirkendall model of radiation-induced segregation to ferritic–martensitic alloys. J. Nucl. Mater. 2012, 425, 117–124. [Google Scholar] [CrossRef]
  32. Wharry, J.P.; Jiao, Z.; Shankar, V.; Busby, J.T.; Was, G.S. Radiation-induced segregation and phase stability in ferritic–martensitic alloy T91. J. Nucl. Mater. 2011, 417, 140–144. [Google Scholar] [CrossRef]
  33. Miller, M.K.; Russell, K.F.; Thompson, K.; Alvis, R.; Larson, D.J. Review of Atom Probe FIB-Based Specimen Preparation Methods. Microsc. Microanal. 2007, 13, 428–436. [Google Scholar] [CrossRef]
  34. London, A.J. Quantifying Uncertainty from Mass-Peak Overlaps in Atom Probe Microscopy. Microsc. Microanal. 2019, 25, 378–388. [Google Scholar] [CrossRef] [Green Version]
  35. Hyde, J.M.; Marquis, E.A.; Wilford, K.B.; Williams, T.J. A sensitivity analysis of the maximum separation method for the characterisation of solute clusters. Ultramicroscopy 2011, 111, 440–447. [Google Scholar] [CrossRef] [PubMed]
  36. Kolli, R.P.; Seidman, D.N. Comparison of Compositional and Morphological Atom-Probe Tomography Analyses for a Multicomponent Fe-Cu Steel. Microsc. Microanal. 2007, 13, 272–284. [Google Scholar] [CrossRef] [Green Version]
  37. Swenson, M.J.; Wharry, J.P. Collected data set size considerations for atom probe cluster analysis. Microsc. Microanal. 2016, 22, 690–691. [Google Scholar] [CrossRef] [Green Version]
  38. Swenson, M.J. The Mechanism of Radiation-Induced Nanocluster Evolution in Oxide Dispersion Strengthened and Ferritic-Martensitic Alloys. Ph.D. Dissertation, Boise State University, Boise, ID, USA, 2017. [Google Scholar]
  39. Williams, C.A.; Haley, D.; Marquis, E.A.; Smith, G.D.W.; Moody, M.P. Defining clusters in APT reconstructions of ODS steels. Ultramicroscopy 2013, 132, 271–278. [Google Scholar] [CrossRef]
  40. Pareige, P.; Miller, M.K.; Stoller, R.E.; Hoelzer, D.T.; Cadel, E.; Radiguet, B. Stability of nanometer-sized oxide clusters in mechanically-alloyed steel under ion-induced displacement cascade damage conditions. J. Nucl. Mater. 2007, 360, 136–142. [Google Scholar] [CrossRef]
  41. Miller, M.K.; Forbes, R.G. Atom-Probe Tomography; Springer: Boston, MA, USA, 2014. [Google Scholar] [CrossRef]
  42. Duh, T.S.; Kai, J.J.; Chen, F.R. Effects of grain boundary misorientation on solute segregation in thermally sensitized and proton-irradiated 304 stainless steel. J. Nucl. Mater. 2000, 283–287, 198–204. [Google Scholar] [CrossRef]
  43. Hu, R.; Smith, G.D.W.; Marquis, E.A. Effect of grain boundary orientation on radiation-induced segregation in a Fe-15.2 at.% Cr alloy. Acta Mater. 2013, 61, 3490–3498. [Google Scholar] [CrossRef]
  44. Hu, C.; Berbenni, S.; Medlin, D.L.; Dingreville, R. Discontinuous segregation patterning across disconnections. Acta Mater. 2023, 246, 118724. [Google Scholar] [CrossRef]
  45. Liu, X.; He, L.; Yan, H.; Bachhav, M.; Stubbins, J.F. A transmission electron microscopy study of EBR-II neutron-irradiated austenitic stainless steel 304 and nickel-base alloy X-750. J. Nucl. Mater. 2020, 528, 151851. [Google Scholar] [CrossRef]
  46. Lach, T.G.; Olszta, M.J.; Taylor, S.D.; Yano, K.H.; Edwards, D.J.; Byun, T.S.; Chou, P.H.; Schreiber, D.K. Correlative STEM-APT characterization of radiation-induced segregation and precipitation of in-service BWR 304 stainless steel. J. Nucl. Mater. 2021, 549, 152894. [Google Scholar] [CrossRef]
  47. Jiao, Z.; Hesterberg, J.; Was, G.S. Effect of post-irradiation annealing on the irradiated microstructure of neutron-irradiated 304L stainless steel. J. Nucl. Mater. 2018, 500, 220–234. [Google Scholar] [CrossRef]
  48. Van Renterghem, W.; Konstantinović, M.J.; Vankeerberghen, M. Evolution of the radiation-induced defect structure in 316 type stainless steel after post-irradiation annealing. J. Nucl. Mater. 2014, 452, 158–165. [Google Scholar] [CrossRef]
  49. Edwards, D.J.; Simonen, E.P.; Garner, F.A.; Greenwood, L.R.; Oliver, B.M.; Bruemmer, S.M. Influence of irradiation temperature and dose gradients on the microstructural evolution in neutron-irradiated 316SS. J. Nucl. Mater. 2003, 317, 32–45. [Google Scholar] [CrossRef]
  50. Maziasz, P.J.; McHargue, C.J. Microstructural evolution in annealed austenitic steels during neutron irradiation. Int. Mater. Rev. 1987, 32, 190–219. [Google Scholar] [CrossRef]
  51. Cawthorne, C.; Brown, C. The occurrence of an ordered FCC phase in neutron irradiated M316 stainless steel. J. Nucl. Mater. 1977, 66, 201–202. [Google Scholar] [CrossRef]
  52. Brager, H.R.; Garner, F.A. Swelling as a consequence of gamma prime (γ’) and M23(C, Si)6 formation in neutron irradiated 316 stainless steel. J. Nucl. Mater. 1978, 73, 9–19. [Google Scholar] [CrossRef]
  53. Fukumoto, K.-I.; Mabuchi, T.; Yabuuchi, K.; Fujii, K. Irradiation hardening of stainless steel model alloy after Fe-ion irradiation and post-irradiation annealing treatment. J. Nucl. Mater. 2021, 557, 153296. [Google Scholar] [CrossRef]
  54. Fukuya, K.; Nakano, M.; Fujii, K.; Torimaru, T.; Kitsunai, Y. Separation of Microstructural and Microchemical Effects in Irradiation Assisted Stress Corrosion Cracking using Post-irradiation Annealing. J. Nucl. Sci. Technol. 2004, 41, 1218–1227. [Google Scholar] [CrossRef]
  55. Zinkle, S.J.; Maziasz, P.J.; Stoller, R.E. Dose dependence of the microstructural evolution in neutron-irradiated austenitic stainless steel. J. Nucl. Mater. 1993, 206, 266–286. [Google Scholar] [CrossRef]
  56. Wang-Koh, Y.M. Understanding the yield behaviour of L12-ordered alloys. Mater. Sci. Technol. 2017, 33, 934–943. [Google Scholar] [CrossRef]
  57. Edwards, D.J.; Simonen, E.P.; Bruemmer, S.M. Evolution of fine-scale defects in stainless steels neutron-irradiated at 275 °C. J. Nucl. Mater. 2003, 317, 13–31. [Google Scholar] [CrossRef]
  58. Kenik, E.A. Radiation-induced segregation in irradiated Type 304 stainless steels. J. Nucl. Mater. 1992, 187, 239–246. [Google Scholar] [CrossRef]
  59. Porter, D.L.; Wood, E.L. In-reactor precipitation and ferritic transformation in neutron-irradiated stainless steels. J. Nucl. Mater. 1979, 83, 90–97. [Google Scholar] [CrossRef] [Green Version]
  60. Kenik, E.A.; Hojou, K. Radiation-induced segregation in FFTF-irradiated austenitic stainless steels. J. Nucl. Mater. 1992, 191–194, 1331–1335. [Google Scholar] [CrossRef]
  61. Was, G.S.; Wharry, J.P.; Frisbie, B.; Wirth, B.D.; Morgan, D.; Tucker, J.D.; Allen, T.R. Assessment of radiation-induced segregation mechanisms in austenitic and ferritic–martensitic alloys. J. Nucl. Mater. 2011, 411, 41–50. [Google Scholar] [CrossRef]
  62. Allen, T.R.; Busby, J.T.; Was, G.S.; Kenik, E.A. On the mechanism of radiation-induced segregation in austenitic Fe–Cr–Ni alloys. J. Nucl. Mater. 1998, 255, 44–58. [Google Scholar] [CrossRef]
  63. Averback, R.S.; Rehn, L.E.; Wagner, W.; Wiedersich, H.; Okamoto, P.R. Kinetics of radiation-induced segregation in Ni-12.7 at.% Si. Phys. Rev. B 1983, 28, 3100–3109. [Google Scholar] [CrossRef]
  64. Batz, W.; Mead, H.W.; Birchenall, C.E. Diffusion of Silicon in Iron. JOM 1952, 4, 1070. [Google Scholar] [CrossRef] [Green Version]
  65. Perkins, R.A. Tracer diffusion of63Ni in Fe-17 wt pct Cr-12 wt pct Ni. Metall. Trans. 1973, 4, 1665–1669. [Google Scholar] [CrossRef]
  66. Arioka, K.; Iijima, Y.; Miyamoto, T. Rapid nickel diffusion in cold-worked type 316 austenitic steel at 360–500 °C. Int. J. Mater. Res. 2017, 108, 791–797. [Google Scholar] [CrossRef]
  67. Anento, N.; Serra, A.; Osetsky, Y. Effect of nickel on point defects diffusion in Fe–Ni alloys. Acta Mater. 2017, 132, 367–373. [Google Scholar] [CrossRef]
  68. Gupta, A.; Kumar, D.; Phatak, V. Asymmetric diffusion at the interfaces in Fe/Si multilayers. Phys. Rev. B 2010, 81, 155402. [Google Scholar] [CrossRef] [Green Version]
  69. Pachaury, Y.; Kumagai, T.; Wharry, J.P.; El-Azab, A. A data science approach for analysis and reconstruction of spinodal-like composition fields in irradiated FeCrAl alloys. Acta Mater. 2022, 234, 118019. [Google Scholar] [CrossRef]
  70. Sen, A.; Bachhav, M.; Pu, X.; Teng, F.; Yao, T.; Wharry, J.P. Irradiation Effects on Stability of δ-UZr2 phase in U-50 wt% Zr Alloy. J. Nucl. Mater. 2023, 576, 154251. [Google Scholar] [CrossRef]
  71. Yao, T.; Sen, A.; Wagner, A.; Teng, F.; Bachhav, M.; EI-Azab, A.; Murray, D.; Gan, J.; Hurley, D.H.; Wharry, J.P.; et al. Understanding spinodal and binodal phase transformations in U-50Zr. Materialia 2021, 16, 101092. [Google Scholar] [CrossRef]
  72. Ke, H.; Wells, P.; Edmondson, P.D.; Almirall, N.; Barnard, L.; Odette, G.R.; Morgan, D. Thermodynamic and kinetic modeling of Mn-Ni-Si precipitates in low-Cu reactor pressure vessel steels. Acta Mater. 2017, 138, 10–26. [Google Scholar] [CrossRef] [Green Version]
  73. Pareige, C.; Kuksenko, V.; Pareige, P. Behaviour of P, Ni impurities and Cr in self ion irradiated Fe–Cr alloys—Comparison to neutron irradiation. J. Nucl. Mater. 2015, 456, 471–476. [Google Scholar] [CrossRef]
  74. Pareige, C.; Etienne, A.; Gueye, P.-M.; Medvedev, A.; Kaden, C.; Konstantinovic, M.J.; Malerba, L. Solute rich cluster formation and Cr precipitation in irradiated Fe-Cr-(Ni,Si,P) alloys: Ion and neutron irradiation. J. Nucl. Mater. 2022, 572, 154060. [Google Scholar] [CrossRef]
  75. Swenson, M.J.; Wharry, J.P. Nanocluster irradiation evolution in Fe-9%Cr ODS and ferritic-martensitic alloys. J. Nucl. Mater. 2017, 496, 24–40. [Google Scholar] [CrossRef]
  76. Swenson, M.J.; Wharry, J.P. The comparison of microstructure and nanocluster evolution in proton and neutron irradiated Fe-9%Cr ODS steel to 3 dpa at 500C. J. Nucl. Mater. 2015, 467, 97–112. [Google Scholar] [CrossRef] [Green Version]
  77. Swenson, M.J.; Wharry, J.P. Rate Theory Model of Irradiation-Induced Solute Clustering in b.c.c. Fe-Based Alloys. JOM 2020, 72, 4017–4027. [Google Scholar] [CrossRef]
  78. Patki, P.V.; Pownell, T.J.; Bazarbayev, Y.; Zhang, D.; Field, K.G.; Wharry, J.P. Systematic study of radiation-induced segregation in neutron-irradiated FeCrAl alloys. J. Nucl. Mater. 2022, 574, 154205. [Google Scholar] [CrossRef]
  79. Tissot, O.; Pareige, C.; Meslin, E.; Décamps, B.; Henry, J. Influence of injected interstitials on α′ precipitation in Fe–Cr alloys under self-ion irradiation. Mater. Res. Lett. 2017, 5, 117–123. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Representative APT needle reconstructions for (a) as-received and (b) proton-irradiated D9, showing absence of clustering of all major alloying species Fe, Cr, Mn, Ni, Si, and Ti.
Figure 1. Representative APT needle reconstructions for (a) as-received and (b) proton-irradiated D9, showing absence of clustering of all major alloying species Fe, Cr, Mn, Ni, Si, and Ti.
Materials 16 04852 g001
Figure 2. Representative APT needle reconstructions for neutron-irradiated D9 showing Ni and Si nanoclustering at (a) 5 dpa, 448 °C, (b) 8 dpa, 430 °C, and (c) 9 dpa, 683 °C; all other major alloying species (Fe, Cr, Mn, Ti) exhibit no evidence of clustering.
Figure 2. Representative APT needle reconstructions for neutron-irradiated D9 showing Ni and Si nanoclustering at (a) 5 dpa, 448 °C, (b) 8 dpa, 430 °C, and (c) 9 dpa, 683 °C; all other major alloying species (Fe, Cr, Mn, Ti) exhibit no evidence of clustering.
Materials 16 04852 g002
Figure 3. Comparison of (a) nanocluster morphologies and compositions and (b) nanocluster size distributions across neutron irradiation conditions.
Figure 3. Comparison of (a) nanocluster morphologies and compositions and (b) nanocluster size distributions across neutron irradiation conditions.
Materials 16 04852 g003
Figure 4. Nanoclusters in (a) 5 dpa, 448 °C, (b) 8 dpa, 430 °C, and (c) 9 dpa, 683 °C D9, showing for each irradiation condition: the Ni and Si distributions in nanoclusters, 10 at% Si isosurfaces, and 30 at% Ni isosurfaces.
Figure 4. Nanoclusters in (a) 5 dpa, 448 °C, (b) 8 dpa, 430 °C, and (c) 9 dpa, 683 °C D9, showing for each irradiation condition: the Ni and Si distributions in nanoclusters, 10 at% Si isosurfaces, and 30 at% Ni isosurfaces.
Materials 16 04852 g004
Figure 5. Example of grain boundary RIS in 8 dpa, 430 °C D9, showing (a) APT needle reconstruction containing all ranged ions and marked region over which composition line scan is taken to obtain RIS profiles of (b) major alloying elements Fe, Ni, and Cr, (c) Mo, Mn, and Ti, and (d) Si.
Figure 5. Example of grain boundary RIS in 8 dpa, 430 °C D9, showing (a) APT needle reconstruction containing all ranged ions and marked region over which composition line scan is taken to obtain RIS profiles of (b) major alloying elements Fe, Ni, and Cr, (c) Mo, Mn, and Ti, and (d) Si.
Materials 16 04852 g005
Figure 6. Bubble plots illustrating the relative (a) size, (b) number density, and (c) Ni/Si ratio of irradiation-induced Ni–Si nanoclusters, nanoprecipitates, and γ’ phases measured in D9 and other austenitic stainless steels in this study (dashed outlines) and from the archival literature (solid lines), specifically refs. [4,7,26,45,46,47,48,49,50,51,52,53,54]; the size of bubble indicates relative size of the parameter being plotted, and the “x” symbols indicate conditions in which no Ni/Si nanoclusters are observed.
Figure 6. Bubble plots illustrating the relative (a) size, (b) number density, and (c) Ni/Si ratio of irradiation-induced Ni–Si nanoclusters, nanoprecipitates, and γ’ phases measured in D9 and other austenitic stainless steels in this study (dashed outlines) and from the archival literature (solid lines), specifically refs. [4,7,26,45,46,47,48,49,50,51,52,53,54]; the size of bubble indicates relative size of the parameter being plotted, and the “x” symbols indicate conditions in which no Ni/Si nanoclusters are observed.
Materials 16 04852 g006
Table 1. Alloy composition (wt%).
Table 1. Alloy composition (wt%).
MethodFeNiCrMoMnSi
Nominal [25]Bal.16.2713.351.551.70
EDS [26]Bal.17.3 ± 0.312.86 ± 0.071.27 ± 0.061.87 ± 0.031.25 ± 0.05
APT65.615.213.952.202.030.66
MethodTiNbVAlCuCo
Nominal [25]0.020.020.01
EDS [26]0.11 ± 0.030.01 ± 0.010.09 ± 0.02
APT0.260.020.020.0140.0050.014
Table 2. Irradiation conditions studied.
Table 2. Irradiation conditions studied.
Irradiating ParticleDose [dpa]Temperature [°C]
Fast neutrons5448 ± 20
Fast neutrons8430 ± 20
Fast neutrons9683 ± 20
2 MeV protons7498 ± 7.5
Table 3. Summary of APT needles from each specimen condition, including APT run parameters, cluster analysis parameters, and number of clusters identified in each needle.
Table 3. Summary of APT needles from each specimen condition, including APT run parameters, cluster analysis parameters, and number of clusters identified in each needle.
Irradiation Condition RHIT File NameLaser Energy [NJ]No. of Ions [million]Pulse Rate [kHz]Detection Rate [%]Base Temp. [K]dmax [nm]Nmin [ions]No. of Clusters
As Received076600.122000.540.3n/an/an/a
077310.19.52500.560.5n/an/an/a
075960.11.22000.555n/an/an/a
Neutron, 5 dpa, 448 °C072730.0852000.549.70.5634141
072740.081.52000.549.70.543139
072790.122001.049.70.583969
075970.12.22000.5550.603270
076620.16.52000.539.50.6041129
075980.14502001.055.10.6040271
Neutron, 8 dpa, 430 °C076050.14352000.854.90.6381351
076060.14482000.5550.6168706
Neutron, 9 dpa, 683 °C072670.072.252000.556.10.492766
072690.192000.548.50.522959
072700.11.32000.749.50.532220
072710.11.52000.549.60.412518
072720.0882000.549.60.502215
Proton, 7 dpa, 500 °C070550.162000.549.7n/an/an/a
070560.15.52000.549.7n/an/an/a
070570.162000.549.6n/an/an/a
070580.060.82000.549n/an/an/a
070680.073.752000.549.7n/an/an/a
Table 4. Average nanocluster morphologies and compositions (at%) determined from APT.
Table 4. Average nanocluster morphologies and compositions (at%) determined from APT.
IrradiationNumber Density [1024 m3]Size [nm]Volume FractionAverage Composition [at%]
FeNiCrMoMnSiTiCNi/Si Ratio
5 dpa
448 °C
2.45 ± 0.995.3 ± 3.80.19 ± 0.0756.5822.5311.360.601.487.000.220.033.22
8 dpa
430 °C
1.24 ± 0.716.8 ± 3.90.21 ± 0.0652.8425.5210.910.841.457.850.280.023.25
9 dpa
683 °C
1.31 ± 0.318.7 ± 8.80.45 ± 0.2345.7132.748.610.481.0410.640.470.023.08
Table 5. Summary of Ni–Si nanocluster, nanoprecipitate, and γ’ phases reported in the archival literature (“n.s.” = not specified, “n.o.” = precipitates not observed, “–” = precipitates present but value not reported).
Table 5. Summary of Ni–Si nanocluster, nanoprecipitate, and γ’ phases reported in the archival literature (“n.s.” = not specified, “n.o.” = precipitates not observed, “–” = precipitates present but value not reported).
AlloySi Conc. [wt%]Irradiating ParticleDose [dpa]Temp. [°C]Average Size [nm]No. Density [1024 m−3]Ni/Si RatioTechniqueRef.
3040.57n (EBR-II)24.15371–389~3516.5TEM[45]
3040.45n3.5288~8–105.5APT[46]
3040.30n (Barsebäck)5.92889.2 ± 0.70.39 ± 0.06APT[47]
3040.78n (Chooz A)243004.7TEM[4]
3040.78n24300100.43TEM[7]
3160.53n (Tihange)803200.005TEM[48]
3160.72n (Tihange)12.234330.06TEM[49]
316n.s.n9.2425–4503.4TEM[50]
316n.s.n (Dounreay)3.2270–5408–230.0008–0.0042~3TEM[51]
316n.s.n (Dounreay)3.2≥540n.o.n.o.n.o.TEM[51]
3160.51n (EBR-II)4852530~0.001–0.0033.14TEM[52]
3160.62n2532040.001TEM[54]
Fe-17Cr-12Ni-1Si0.966.4 MeV Fe3+2–6.72002.51.8–22.1–2.5APT[53]
D90.66Neutron5448<13TEM[26]
D90.66Neutron8.24303.5 ± 0.10.008 ± 0.0033TEM[26]
D90.66Neutron9.26835.1 ± 0.30.002 ± 0.001<3TEM[26]
D90.66Neutron54485.27 ± 3.812.45 ± 0.993.22 ± 0.87APTThis study
D90.66Neutron84306.82 ± 3.881.24 ± 0.713.25 ± 0.74APTThis study
D90.66Neutron96838.70 ± 8.811.31 ± 0.313.08 ± 1.24APTThis study
D90.66Proton7500n.o.n.o.n.o.APTThis study
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mullurkara, S.V.; Bejawada, A.; Sen, A.; Sun, C.; Bachhav, M.; Wharry, J.P. Nanocluster Evolution in D9 Austenitic Steel under Neutron and Proton Irradiation. Materials 2023, 16, 4852. https://doi.org/10.3390/ma16134852

AMA Style

Mullurkara SV, Bejawada A, Sen A, Sun C, Bachhav M, Wharry JP. Nanocluster Evolution in D9 Austenitic Steel under Neutron and Proton Irradiation. Materials. 2023; 16(13):4852. https://doi.org/10.3390/ma16134852

Chicago/Turabian Style

Mullurkara, Suraj Venkateshwaran, Akshara Bejawada, Amrita Sen, Cheng Sun, Mukesh Bachhav, and Janelle P. Wharry. 2023. "Nanocluster Evolution in D9 Austenitic Steel under Neutron and Proton Irradiation" Materials 16, no. 13: 4852. https://doi.org/10.3390/ma16134852

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop