Effect of alloying elements on the microstructure and mechanical properties of nanostructured...

6
Letter Effect of alloying elements on the microstructure and mechanical properties of nanostructured ferritic steels produced by spark plasma sintering Somayeh Pasebani, Indrajit Charit Department of Chemical and Materials Engineering, University of Idaho, Moscow, ID 83844, USA article info Article history: Received 23 November 2013 Received in revised form 23 January 2014 Accepted 29 January 2014 Available online 15 February 2014 Keywords: Nanostructures Mechanical alloying Powder metallurgy Transmission electron microscopy High temperature alloy abstract Several Fe–14Cr based alloys with varying compositions were processed using a combined route of mechanical alloying and spark plasma sintering. Microstructural characteristics of the consolidated alloys were examined via transmission electron microscopy and atom probe tomography, and mechanical prop- erties evaluated using microhardness testing. Lanthanum oxide (0.5 wt.%) was added to Fe–14Cr leading to improvement in microstructural stability and mechanical properties mainly due to a high number den- sity of La–Cr–O-enriched nanoclusters. The combined addition of La, Ti (1 wt.%) and Mo (0.3 wt.%) to the Fe–14Cr base composition further enhanced the microstructural stability and mechanical properties. Nanoclusters enriched in Cr–Ti–La–O with a number density of 1.4 10 24 m 3 were found in this alloy with a bimodal grain size distribution. After adding Y 2 O 3 (0.3 wt.%) along with Ti and Mo to the Fe– 14Cr matrix, a high number density (1.5 10 24 m 3 ) of Cr–Ti–Y–O-enriched NCs was also detected. For- mation mechanism of these nanoclusters can be explained through the concentrations and diffusion rates of the initial oxide species formed during the milling process and initial stages of sintering as well as the thermodynamic nucleation barrier and their enthalpy of formation. Ó 2014 Elsevier B.V. All rights reserved. 1. Introduction Nanostructured ferritic steels (NFSs), a subcategory of oxide dis- persion strengthened (ODS) steels, have outstanding high temper- ature strength, creep strength [1,2] and excellent radiation damage resistance [3]. These enhanced properties of NFSs have been attrib- uted to the high number density of Y–Ti–O-enriched nanoclusters (NCs) with diameter of 1–2 nm [4]. The Y–Ti–O-enriched NCs have been found to be stable under irradiation and effective in trapping helium [5]. These NCs are formed due to the mechanical alloying (MA) of Fe–Cr–Ti powder with Y 2 O 3 during high energy ball milling followed by hot consolidation route such as hot isostatic pressing (HIP) or hot extrusion [6–8]. Alinger et al. [4] have investigated the effect of alloying elements on the formation mechanism of NCs in NFSs processed by hot isostatic pressing (HIP) and reported both Ti and high milling energy were necessary for the formation of NCs. Miller and Parish [9] suggested that the excellent creep properties in yttria-bearing NFSs result from the pinning of the grain boundaries by a combined effect of solute segregation and precipitation. Although HIP and hot extrusion are commonly used to consoli- date the NFSs, anisotropic properties and processing costs are con- sidered challenging issues. Recently, spark plasma sintering (SPS) has been utilized to sinter the powder at a higher heating rate, low- er temperature and shorter dwell time. This can be done by apply- ing a uniaxial pressure and direct current pulses simultaneously to a powder sample contained in a graphite die [10]. Except for a few studies on consolidation of simple systems such as Fe–9Cr–0.3/ 0.6Y 2 O 3 [11] and Fe–14Cr–0.3Y 2 O 3 [10], the SPS process has not been extensively utilized to consolidate the NFSs with complex compositions. Recently, the role of Ti and Y 2 O 3 in processing of Fe–16Cr–3Al–1Ti–0.5Y 2 O 3 (wt.%) via MA and SPS was investigated by Allahar et al. [12]. A bimodal grain size distribution in conjunc- tion with Y–Ti–O-enriched NCs were obtained [12,13]. In this study, Fe–14Cr (wt.%) was designed as the base or matrix alloy, and then Ti, La 2 O 3 and Mo were sequentially added to the ferritic matrix and ball milled. This approach allowed us to study the effect of individual and combined addition of solutes on the formation of NCs along with other microstructural evolutions. Furthermore, SPS instead of other traditional consolidation methods was used to consolidate the NFS powder. The mixture http://dx.doi.org/10.1016/j.jallcom.2014.01.243 0925-8388/Ó 2014 Elsevier B.V. All rights reserved. Corresponding author. Tel.: +1 208 885 5964; fax: +1 208 885 7462. E-mail address: [email protected] (I. Charit). Journal of Alloys and Compounds 599 (2014) 206–211 Contents lists available at ScienceDirect Journal of Alloys and Compounds journal homepage: www.elsevier.com/locate/jalcom

Transcript of Effect of alloying elements on the microstructure and mechanical properties of nanostructured...

Page 1: Effect of alloying elements on the microstructure and mechanical properties of nanostructured ferritic steels produced by spark plasma sintering

Journal of Alloys and Compounds 599 (2014) 206–211

Contents lists available at ScienceDirect

Journal of Alloys and Compounds

journal homepage: www.elsevier .com/locate / ja lcom

Letter

Effect of alloying elements on the microstructure and mechanicalproperties of nanostructured ferritic steels produced by sparkplasma sintering

http://dx.doi.org/10.1016/j.jallcom.2014.01.2430925-8388/� 2014 Elsevier B.V. All rights reserved.

⇑ Corresponding author. Tel.: +1 208 885 5964; fax: +1 208 885 7462.E-mail address: [email protected] (I. Charit).

Somayeh Pasebani, Indrajit Charit ⇑Department of Chemical and Materials Engineering, University of Idaho, Moscow, ID 83844, USA

a r t i c l e i n f o

Article history:Received 23 November 2013Received in revised form 23 January 2014Accepted 29 January 2014Available online 15 February 2014

Keywords:NanostructuresMechanical alloyingPowder metallurgyTransmission electron microscopyHigh temperature alloy

a b s t r a c t

Several Fe–14Cr based alloys with varying compositions were processed using a combined route ofmechanical alloying and spark plasma sintering. Microstructural characteristics of the consolidated alloyswere examined via transmission electron microscopy and atom probe tomography, and mechanical prop-erties evaluated using microhardness testing. Lanthanum oxide (0.5 wt.%) was added to Fe–14Cr leadingto improvement in microstructural stability and mechanical properties mainly due to a high number den-sity of La–Cr–O-enriched nanoclusters. The combined addition of La, Ti (1 wt.%) and Mo (0.3 wt.%) to theFe–14Cr base composition further enhanced the microstructural stability and mechanical properties.Nanoclusters enriched in Cr–Ti–La–O with a number density of 1.4 � 1024 m�3 were found in this alloywith a bimodal grain size distribution. After adding Y2O3 (0.3 wt.%) along with Ti and Mo to the Fe–14Cr matrix, a high number density (1.5 � 1024 m�3) of Cr–Ti–Y–O-enriched NCs was also detected. For-mation mechanism of these nanoclusters can be explained through the concentrations and diffusion ratesof the initial oxide species formed during the milling process and initial stages of sintering as well as thethermodynamic nucleation barrier and their enthalpy of formation.

� 2014 Elsevier B.V. All rights reserved.

1. Introduction

Nanostructured ferritic steels (NFSs), a subcategory of oxide dis-persion strengthened (ODS) steels, have outstanding high temper-ature strength, creep strength [1,2] and excellent radiation damageresistance [3]. These enhanced properties of NFSs have been attrib-uted to the high number density of Y–Ti–O-enriched nanoclusters(NCs) with diameter of 1–2 nm [4]. The Y–Ti–O-enriched NCs havebeen found to be stable under irradiation and effective in trappinghelium [5]. These NCs are formed due to the mechanical alloying(MA) of Fe–Cr–Ti powder with Y2O3 during high energy ball millingfollowed by hot consolidation route such as hot isostatic pressing(HIP) or hot extrusion [6–8]. Alinger et al. [4] have investigatedthe effect of alloying elements on the formation mechanism ofNCs in NFSs processed by hot isostatic pressing (HIP) and reportedboth Ti and high milling energy were necessary for the formationof NCs. Miller and Parish [9] suggested that the excellent creepproperties in yttria-bearing NFSs result from the pinning of the

grain boundaries by a combined effect of solute segregation andprecipitation.

Although HIP and hot extrusion are commonly used to consoli-date the NFSs, anisotropic properties and processing costs are con-sidered challenging issues. Recently, spark plasma sintering (SPS)has been utilized to sinter the powder at a higher heating rate, low-er temperature and shorter dwell time. This can be done by apply-ing a uniaxial pressure and direct current pulses simultaneously toa powder sample contained in a graphite die [10]. Except for a fewstudies on consolidation of simple systems such as Fe–9Cr–0.3/0.6Y2O3 [11] and Fe–14Cr–0.3Y2O3 [10], the SPS process has notbeen extensively utilized to consolidate the NFSs with complexcompositions. Recently, the role of Ti and Y2O3 in processing ofFe–16Cr–3Al–1Ti–0.5Y2O3 (wt.%) via MA and SPS was investigatedby Allahar et al. [12]. A bimodal grain size distribution in conjunc-tion with Y–Ti–O-enriched NCs were obtained [12,13].

In this study, Fe–14Cr (wt.%) was designed as the base or matrixalloy, and then Ti, La2O3 and Mo were sequentially added to theferritic matrix and ball milled. This approach allowed us to studythe effect of individual and combined addition of solutes on theformation of NCs along with other microstructural evolutions.Furthermore, SPS instead of other traditional consolidationmethods was used to consolidate the NFS powder. The mixture

Page 2: Effect of alloying elements on the microstructure and mechanical properties of nanostructured ferritic steels produced by spark plasma sintering

S. Pasebani, I. Charit / Journal of Alloys and Compounds 599 (2014) 206–211 207

of Fe–Cr–Ti–Mo powder with Y2O3 was also processed andcharacterized in a similar manner for comparison with the rest ofthe developed alloys.

2. Experimental

The chemical compositions of all the developed alloys along with their identi-fying names in this study are given in Table 1. High energy ball milling was per-formed in a SPEX 8000M shaker mill for 10 h using Ar atmosphere with themilling media as steel balls of8 mm in diameter and a ball to powder ratio (BPR)of 10:1. A Dr. Sinter Lab SPS-515S was used to consolidate the as-milled powderat different temperatures (850, 950 and 1050 �C) for 7 min using the pulse pattern12–2 ms, a heating rate of 100 �C/min and a pressure of 80 MPa. The SPSed sampleswere in the form of disks with 8 mm in height and 12 mm in diameter.

The density of the sintered specimens was measured by Archimedes’ method.Vickers microhardness tests were performed using a Leco LM100 microhardnesstester operated at a load of 1000 g–f (9.8 N). A Fischione Model 110 Twin-Jet Elec-tropolisher containing a mixture of CH3OH–HNO3 (80:20 by vol.%) as the electrolyteand operated at about �40 �C was used to prepare specimens for transmission elec-tron microscopy (TEM). A FEI Tecnai TF30–FEG STEM operating at 300 kV was used.The energy dispersive spectroscopy (EDS) attached with the STEM was used toroughly examine the chemical composition of the particles. A Quanta 3D FEGinstrument with a Ga-ion source focused ion beam (FIB) was used to prepare spec-imens for atom probe tomography (APT) studies on 14L, 14LMT and 14YMT sam-ples. The APT analysis was carried out using a CAMECA LEAP 4000X HRinstrument operating in the voltage mode at 50–60 K and 20% of the standing volt-age pulse fraction. The atom maps were reconstructed using CAMECA IVAS 3.6 soft-ware and the maximum separation algorithm to estimate the size and chemicalcomposition of NCs. This was applied to APT datasets each containing 20–30 millionions for each specimen. Lower evaporation field of the nanoparticles and trajectoryaberrations caused estimation of higher Fe atoms in the nanoclusters. Although thecontribution of Fe atoms from the matrix was examined here, the matrix-correctionwas not addressed in this study.

3. Results and discussion

The TEM bright field micrographs for the various alloys SPSed at950 �C for 7 min are illustrated in Fig. 1a–d. The microstructure of14Cr alloy shown in Fig. 1a revealed a complex microstructurewith submicron subgrain-like structures, relatively high densityof dislocations and low number density of oxide nanoparticles.The nanoparticles were larger (25–65 nm) than the other SPSed al-loys and found to have chemical compositions close to Cr2O3 andFeCr2O4 as analyzed by energy dispersive spectroscopy.

The microstructure of the consolidated 14L alloy is shown inFig. 1b. The microstructure consisted of more ultra fine grains(<1 lm but >100 nm), a few nanograins with sharp boundariesand a higher number of nanoparticles mainly in the grain interiors.The number density of nanoparticles was higher than that of 14Cralloy shown in Fig. 1a but lower than 14LMT (Fig. 1c) and 14YMT(Fig. 1d). In 14L alloy, the nanoparticles with 2–11 nm in diameterwere found inside the grains (hard to be observed at magnificationgiven in Fig. 1b and micrographs taken at higher magnificationswas used for this purpose) whereas the nanoparticles with

Table 1The alloy compositions and processing conditions (milled for 10 h and SPSed at 850-1050 �C for 7 min).

Alloy ID Elements (wt.%)

Cr Ti La2O3 Y2O3 Mo Fe

14Cr-unmilled 14 0 0 0 0 Bal.14Cr 14 0 0 0 0 Bal.14T 14 1 0 0 0 Bal.14L 14 0 0.5 0 0 Bal.14Y 14 0 0 0.3 0 Bal.14LM 14 0 0.5 0 0.3 Bal.14LT 14 1 0.5 0 0 Bal.14LMT(0.3) 14 1 0.3 0 0.3 Bal.14LMT 14 1 0.5 0 0.3 Bal.14LMT(0.7) 14 1 0.7 0 0.3 Bal.14YMT 14 1 0 0.3 0.3 Bal.

50–80 nm in diameter were located at the grain boundary regions.The particles on the boundaries are likely to be mainly Cr2O3 andLaCrO3, but the chemical analysis of those smallest particles couldnot be done precisely due to the significant influence of the ferriticmatrix.

Fig. 1c shows the microstructure of the SPSed 14LMT alloy,consisting of both ultra fine grains (as defined previously) andnanograins (6100 nm). The nanoparticles present in the micro-structure were complex oxides of Fe, Cr and Ti. The nanoparticleswith faceted morphology and smaller than 10 nm in diameter wereenriched in La and Ti. No evidence of stoichiometric La2TiO5 orLa2Ti2O7 particles was observed based on the EDS and diffractiondata. A similar type of microstructure was revealed in the SPSed14YMT alloy as shown in Fig. 1d.

The particle size distribution histograms of the 14Cr, 14L,14LMT and 14YMT alloys are plotted in Fig. 2a–d, respectively.Approximately 1000 particles were sampled from each alloy to de-velop the histograms. The average particle size decreased in orderof 14Cr, 14L, 14LMT and 14YMT. The highest fraction of the particlesize as shown in the histograms of 14Cr, 14L, 14LMT and 14YMTwas found to be associated with 25 ± 5 nm (18 ± 2.5%), 10 ± 5 nm(28 ± 3%), 5 ± 1 nm (40 ± 6%) and 5 ± 1 nm (46 ± 5%) in diameter,respectively. The number density of nanoparticles smaller than5 ± 1 nm was higher in 14YMT than 14LMT alloy.

The 3-D APT maps for 14L alloy revealed a number density(�3 � 1022 m�3) of CrO–La–O-enriched NCs. The average Guinierradius of these NCs was 1.9 ± 0.6 nm. The average composition ofthe NCs in 14L was estimated by using the maximum separationalgorithm to be Fe–17.87 ± 3.4Cr–32.61 ± 3.2O–8.21 ± 1.1La (at.%).A higher number density (�1.4 � 1024 m�3) and smaller NCs withaverage Guinier radius of 1.43 ± 0.20 nm were observed in theAPT maps for 14LMT alloy as shown in Fig. 3a. The NCs were Cr–Ti–La–O-enriched with the average composition of Fe–10.9 ± 2.8Cr–30.9 ± 3.1O–17.3 ± 2.5Ti–8.2 ± 2.2La (at.%). Accordingto the LEAP measurements, the chemical composition of NCs dif-fered considerably from stoichiometric oxides. A large amount ofFe and Cr was detected inside the NCs, and La/Ti and La/O ratioswere not consistent with La2TiO5 or La2Ti2O7 as expected basedon thermodynamic calculations, rather the ratios were sub-stoichi-ometric. The 3-D APT maps for 14YMT alloy were similar to 14YMTalloy as shown in Fig. 3b. The NCs with an average radius of1.24 ± 0.2 nm and a number density of 1.5 � 1024 m�3 were Cr–Ti–Y–O-enriched. The chemical composition of NCs was estimatedclose to Fe–8.52 ± 3.1Cr–37.39 ± 4.5O–24.52 ± 3.1Ti–10.95 ± 3.1Y(at.%). The matrix-corrected compositions are currently being ana-lyzed and will be reported in a full-length publication in nearfuture.

The relative density of various alloys sintered at 850–1050 �C isshown in Fig. 4a. Generally, a higher density was obtained in thespecimens sintered at higher temperatures. At 850 and 950 �C,the density of unmilled 14Cr specimen (97.2% and 97.5%) washigher than the milled/SPSed 14Cr (92.8% and 95.5%) because theunmilled powder particles were less hard (due to absence of strainhardening) and plastically deformed to a higher degree than themilled powder leading to a higher density. Adding 0.5 and0.7 wt.% of La2O3 and 0.3 wt.% Y2O3 to the 14Cr matrix significantlydecreased the density of the specimen, especially at 850 and950 �C; however, adding Ti to 14L and 14Y improved the densityto some extent.

The microhardness data of various alloys processed at differenttemperatures are shown in Fig. 4b. In general, microhardness in-creased with increasing SPS temperatures up to 950 �C and thendecreased. Both Y and La increased the hardness due to the disper-sion hardening effect. The hardness increased at the higher contentof La due to the greater effect of dispersion hardening. Adding Tiseparately to the 14Cr matrix improved the hardness due to the

Page 3: Effect of alloying elements on the microstructure and mechanical properties of nanostructured ferritic steels produced by spark plasma sintering

Fig. 2. Particle size frequency histogram for (a) 14Cr, (b) 14L, (c) 14LMT and (d) 14YMT alloys.

Fig. 1. TEM bright field micrographs for various alloys (a) 14Cr, (b) 14L, (c) 14LMT and (d) 14YMT.

208 S. Pasebani, I. Charit / Journal of Alloys and Compounds 599 (2014) 206–211

dispersion hardening but only at lower temperature (850 �C). Thecoarsening of Ti-enriched particles at above 850 �C plausibly

decreased the hardness. However, at 950 �C, higher hardness(457 HV) was achieved by a combined addition of La and Ti to

Page 4: Effect of alloying elements on the microstructure and mechanical properties of nanostructured ferritic steels produced by spark plasma sintering

Fig. 3. Three-dimensional atom maps showing NCs for (a) 14LMT – 91 � 34 � 30 nm3 and (b) 14YMT – 93 � 30 � 30 nm3.

Fig. 4. (a) The relative density and (b) microhardness values for different SPSed alloys processed at different SPS temperatures for a dwell time of 7 min.

S. Pasebani, I. Charit / Journal of Alloys and Compounds 599 (2014) 206–211 209

the 14Cr matrix to produce 14LT. Further addition of Mo to 14LTimproved the hardness through solid solution strengthening in14LMT (495 HV).

High dislocation density and no well-defined grain boundarieswere characteristics of 14Cr alloy as shown in Fig. 1a. The presenceof a low number density and larger oxide particles (FeCr2O4 andCr2O3) at the boundaries could not create an effective pinningeffect during sintering. As a result, some of these particles becameconfined within the grain interiors. The coarse grains had thecapacity to produce and store high density of dislocations that

subsequently resulted in the strain hardening effect. The hardeningmechanism in 14Cr alloy can thus be attributed to greater disloca-tion activities and resulting strain hardening effect. The grainboundary or precipitation hardening cannot be the dominantmechanism because of larger particles, greater inter-particle spac-ing and weakened Zener drag effect at the temperature of sinter-ing. Such strain hardening capability in nanocrystalline Feconsolidated via SPS was reported by other researchers, too[14,15]. Interestingly, the high hardness in Fe–14Cr alloy consoli-dated via SPS at 1100 �C for 4 min by Auger et al. [10] was

Page 5: Effect of alloying elements on the microstructure and mechanical properties of nanostructured ferritic steels produced by spark plasma sintering

210 S. Pasebani, I. Charit / Journal of Alloys and Compounds 599 (2014) 206–211

attributed to the formation of martensitic laths caused by highercarbon content diffusing from the die, possible Cr segregationand rapid cooling during SPS. It is noteworthy to mention that nomartensite lath was observed in the consolidated 14Cr alloy inthe present study.

The level of solutes in the bcc matrix could be much greaterthan the equilibrium level, associated with a large number ofvacancies created during milling. Our recent study [16] has shownthat high energy ball milling has a complex role in initiating nucle-ation of La–Ti–O-enriched NCs in 14LMT alloy powder, with amean radius of �1 nm, a number density of 3.7 � 10�24 m�3 anda composition of Fe–12.11Cr–9.07Ti–4.05La (at.%). The initiationof NCs during ball milling of NFSs has also been investigated byother researchers [8,17,18]. According to Williams et al. [8], dueto a low equilibrium solubility of O in the matrix, the precipitationof nanoparticles is driven by an oxidation reaction, subsequentlyresulting in reduction of the free energy. As the SPS proceeds, thenumber density of NCs would decrease and larger grain boundaryoxides would form with the grain structure developing simulta-neously during the sintering process [8]. Formation of larger grainboundary oxides as shown in Fig. 1a could have been preceded bysegregation of O and Cr to grain boundaries leading to a decrease inthe level of the solutes in the ferritic matrix. The initial oxidesforming in a chromium-rich matrix can be Cr2O3 as suggested byWilliams et al. [8]. However, formation of LaCrO3 in 14L alloy(shown in Fig. 1b) was associated with a higher reduction in thefree energy according to the enthalpy of formation of various oxi-des given in Table 2. The presence of nanoparticles caused grainboundary pinning and subsequently stabilized the nanocrystallinegrains.

The high density of defects (dislocations and vacancies) in asupersaturated solid solution, such as 14LMT and 14YMT alloys,could dramatically increase the driving force for accelerated sub-grain formation during the initial stage of sintering. At the initialstage, the vacancies created during the milling are annihilated[8,17]. Meanwhile, the temperature is not high enough to producea significant number of thermal vacancies; subsequently, anynucleation of new NCs will be prevented. As the SPS proceeds withno nucleation of new NCs, the high concentrations of extra solutesin the matrix are thermodynamically and kinetically required toprecipitate out to form larger oxide particles. The larger solute-enriched oxide particles can be formed more favorably on the grainboundaries due to the higher boundary diffusivity. On the otherhand, it should be considered that there is a dynamic plasticdeformation occurring within the powder particles during SPS.The interaction of larger particles and dislocations introduced by

Table 2The standard enthalpies of formation of various oxide compounds at 25 �C [8,21,22].

Element Composition �DHf (kJ mol�1(oxide))

Cr Cr2O3 1131CrO2 583

Fe Fe3O4 1118Fe2O3 822

Ti TiO 543TiO2 944Ti2O3 1522Ti3O5 2475

Y Y2O3 1907YCrO3 1493Y2Ti2O7 3874

La La2O3 1794La2Ti2O7 3855LaCrO3 1536

dynamic hot deformation can explain the coarsening in somegrains; because larger particles could not effectively pin the dislo-cations and the grain boundary migration could be facilitated fol-lowing the orientation with lower efficiency of Zener dragmechanism [19]. Once the extra solutes present in the matrix pre-cipitated out, the microstructure will remain very stable because ofthe grain boundary pinning by triple-junctions of the grain bound-aries themselves [20], along with the high density of NCs and otherultrafine oxide particles [8]. Further coarsening of the grains willbe prevented even for longer dwell times at 950 �C. Therefore, a bi-modal grain size distribution emerged. The hardening of 14LMTand 14YMT alloys were attributed to a combined effect of solidsolution strengthening, Hall-Petch strengthening and precipitationhardening.

Based on the APT studies of the as-milled powder [16] and for-mation mechanism of the oxide particles suggested by Williamset al. [8] it could be speculated that in 14LMT and 14YMT alloys,Cr–O species form first and then absorb Ti and La/Y. This is associ-ated with a change in the interfacial energy of Cr–O species eventhough it is not thermodynamically the most favorable oxide. Ithas been established that the driving force for the oxide precipi-tates to form is the low solubility limit of oxygen in the ferritic ma-trix. The change in free energy due to oxidation reaction andnucleation of oxide nanoparticles is the leading mechanism [8].

The majority of the oxygen required to generate the oxide nano-particles may be provided from the surface oxide during millingprocess. Furthermore, higher concentrations of Cr led to greaternucleation of Cr–O by influencing the kinetics of oxide formation.Concentrations and diffusivities of the oxide species along withthe energy barrier for nucleation will control the nucleation ofoxide nanoparticles. After the Cr–O formed during sintering, theTi–O and Y/La-enriched clusters could form. The sub-stoichiome-tric NCs in 14LMT and 14YMT alloys were not due to insufficientlevel of O in the matrix [8]. Formation of stoichiometric Y2Ti2O7

and Y2TiO5 requires very high temperatures [8], which wereoutside the scope of this study.

4. Conclusion

The SPSed Fe–14Cr alloy was found to have a higher hardness atroom temperature due to the strain hardening effect. The stabilityof its microstructure at high temperatures was improved by addi-tion of La forming the Cr–La–O-enriched NCs. Adding La and Ti toFe–14Cr matrix significantly improved the mechanical behaviorand microstructural stability further due to the high numberdensity of Cr–Ti–La–O-enriched NCs in 14LMT alloy. It is demon-strated that the potential capability of La in developing new NFSsis promising but further investigations on their thermal andirradiation stability will still be required.

Acknowledgement

This work was supported partly by the Laboratory DirectedResearch and Development Program of Idaho National Laboratory(INL), and by the Advanced Test Reactor National Scientific UserFacility (ATR NSUF), Contract DE-AC07-05ID14517. The authorsgratefully acknowledge the assistance of the staff members at theMicroscopy and Characterization Suite (MaCS) facility at the Centerfor Advanced Energy Studies (CAES).

References

[1] M.J. Alinger, G.R. Odette, G.E. Lucas, J. Nucl. Mater. 307–311 (2002) 484.[2] R.L. Klueh, J.P. Shingledecker, R.W. Swindeman, D.T. Hoelzer, J. Nucl. Mater.

341 (2005) 103.[3] M.J. Alinger, G.R. Odette, D.T. Hoelzer, J. Nucl. Mater. 329–333 (2004) 382.[4] M.J. Alinger, G.R. Odette, D.T. Hoelzer, Acta Mater. 57 (2009) 392.

Page 6: Effect of alloying elements on the microstructure and mechanical properties of nanostructured ferritic steels produced by spark plasma sintering

S. Pasebani, I. Charit / Journal of Alloys and Compounds 599 (2014) 206–211 211

[5] G.R. Odette, M.L. Alinger, B.D. Wirth, Annu. Rev. Mater. Res. 38 (2008) 471.[6] S. Ukai, T. Okuda, M. Fujiwara, T. Kobayashi, S. Mizuta, H. Nakashima, J. Nucl.

Sci. Technol. 39 (2002) 872.[7] S. Ukai, M. Fujiwara, J. Nucl. Mater. 307–311 (2002) 749.[8] C.A. Williams, P. Unifantowicz, N. Baluc, G.D. Smith, E.A. Marquis, Acta Mater.

61 (2013) 2219.[9] M.K. Miller, C.M. Parish, Mater. Sci. Technol. 27 (2011) 729.

[10] M.A. Auger, V. De Castro, T. Leguey, A. Muñoz, Pareja, R, J. Nucl. Mater. 436(2013) 68.

[11] C. Heintze, M. Hernández-Mayoral, A. Ulbricht, F. Bergner, A. Shariq, T.Weissgärber, H. Frielinghaus, J. Nucl. Mater. 428 (2012) 139.

[12] K.N. Allahar, J. Burns, B. Jaques, Y.Q. Wu, I. Charit, J.I. Cole, D.P. Butt, J. Nucl.Mater. 443 (2013) 256.

[13] Y.Q. Wu, K.N. Allahar, J. Burns, B. Jaques, I. Charit, D.P. Butt, J.I. Cole, Cryst. Res.Technol. (2013) 1, http://dx.doi.org/10.1002/crat.201300173.

[14] K. Oh-Ishi, H.W. Zhang Hw, T. Ohkubo, K. Hono, Mater. Sci. Eng. A 456 (2007)20.

[15] B. Srinivasarao, K. Ohishi, T. Ohkubo, K. Hono, Acta Mater. 57 (2009) 3277.[16] S. Pasebani, I. Charit, Y.Q. Wu, D.P. Butt, J.I. Cole, Acta Mater. 61 (2013)

5605.[17] M.L. Brocq, F. Legendre, M.H. Mathon, A. Mascaro, S. Poissonnet, B. Radiguet, P.

Pareige, M. Loyer, O. Leseigneur, Acta Mater. 60 (2012) 7150.[18] M. Brocq, B. Radiguet, S. Poissonnet, F. Cuvilly, P. Pareige, F. Legendre, J. Nucl.

Mater. 409 (2011) 80.[19] H.K.D.H. Bhadeshia, Mater. Sci. Eng. A 223 (1997) 64.[20] H.K.D.H. Bhadeshia, Mater. Sci. Technol. 16 (2000) 1404.[21] W. Gale, T. Totemeier, Smithells Metals Reference Book, Amsterdam, Holland,

2004.[22] T.J. Kallarackel, S. Gupta, P. Singh, J. Am. Ceram. Soc. (2013) 1, http://

dx.doi.org/10.1111/jace.12435.