You are on page 1of 45

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/321094461

Effects of Laves phase particles on recovery and recrystallization behaviors of


Nb-containing FeCrAl alloys

Article  in  Acta Materialia · November 2017


DOI: 10.1016/j.actamat.2017.11.027

CITATIONS READS

3 105

3 authors, including:

Zhiqian Sun
Oak Ridge National Laboratory
16 PUBLICATIONS   162 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

High Temperature Steel for Ultra-supercritical steam turbine View project

All content following this page was uploaded by Zhiqian Sun on 10 January 2018.

The user has requested enhancement of the downloaded file.


Notice: This manuscript has been authored by UT-Battelle, LLC, under contract DE-AC05-00OR22725 with the

US Department of Energy (DOE). The US government retains and the publisher, by accepting the article for

publication, acknowledges that the US government retains a nonexclusive, paid-up, irrevocable, worldwide

license to publish or reproduce the published form of this manuscript, or allow others to do so, for US government

purposes. DOE will provide public access to these results of federally sponsored research in accordance with the

DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan).

1
Effects of Laves phase particles on recovery and recrystallization behaviors of Nb-containing

FeCrAl alloys

Zhiqian Sun*1, Philip D. Edmondson1, and Yukinori Yamamoto1

1. Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA

ABSTRACT

The microstructures and mechanical properties of deformed and annealed Nb-containing FeCrAl alloys were

investigated. Fine dispersion of Fe2Nb-type Laves phase particles was observed in the bcc-Fe matrix after

applying a thermomechanical treatment, especially along grain/subgrain boundaries, which effectively stabilized

the recovered and recrystallized microstructures compared with the Nb-free FeCrAl alloy. The stability of

recovered areas increased with Nb content up to 1 wt%. The recrystallized grain structure in Nb-containing

FeCrAl alloys consisted of elongated grains along the rolling direction with a weak texture when annealed below

1100 °C. An abnormal relationship between recrystallized grain size and annealing temperature was found.

Microstructural inhomogeneity in the deformed and annealed states was explained based on the Taylor factor.

Annealed Nb-containing FeCrAl alloys showed a good combination of strength and ductility, which is desirable

for their application as fuel cladding in light-water reactors.

Keywords: FeCrAl alloys; Recrystallization; Laves phases; Texture; EBSD

* Corresponding author.

Address: Oak Ridge National Laboratory, PO Box 2008, MS6115, Oak Ridge, TN 37831, USA

Email: sunz@ornl.gov

Tel: 01-865-208-7295

2
1. INTRODUCTION

Iron-chromium-aluminum (FeCrAl) ferritic alloys have long been used as heat-resistant materials for many

industrial applications. Currently, wrought FeCrAl alloys are under development as promising accident-tolerant

fuel cladding for light water reactors (LWRs) to provide larger safety margins [1-6]. In accident scenarios (e.g.,

loss-of-coolant accidents), the fuel temperature quickly increases because of the decay heat production and drop

in heat conductance. Commercial zirconium alloy fuel cladding is susceptible to severe degradation in high-

temperature (>1200 °C) steam environments [7, 8]. Unlike zirconium alloys, FeCrAl alloys demonstrate superior

oxidation resistance at high temperatures through the formation of continuous external alumina scales [9, 10]. The

amounts of Cr (~10-13 wt%) and Al (~5-6 wt%) in FeCrAl alloys have been tuned to provide improved surface

protection effects at both the service temperature and elevated temperatures in accident scenarios. Minor alloying

additions have also been made. The additions of reactive elements, such as yttrium (Y), benefit the adhesion

between the substrate matrix of the alloys and protective external aluminum-oxide scales [9]; and molybdenum

(Mo) and niobium (Nb) have been employed to strengthen the body-centered cubic (bcc)-Fe matrix through solid-

solution hardening and control microstructures [2, 11, 12]. Improving the processability of FeCrAl alloys is also

an important target for the production of seamless thin-wall FeCrAl alloy tubes as fuel cladding, so optimization

of production process parameters has been pursued [12-15]. The goal of nuclear-grade FeCrAl alloy development

is to build a mature combination of alloy compositions and optimized process pathways for fabricating seamless

thin-wall FeCrAl tube products that are suitable for the mass production of FeCrAl alloy fuel cladding and

contribute to the development of accident-tolerant fuels.

There are several factors limiting the processability of FeCrAl alloys. First, because of the high Al and Cr content,

FeCrAl alloys have a bcc structure up to the melting temperature [16], eliminating the chance of grain refinement

by phase transformation that occurs in carbon steels. A similar characteristic is observed in several ferrous alloys,

including high-Cr ferritic stainless steels, low-density Fe-Al alloys [17, 18], and NiAl/Ni2TiAl-strengthened

ferritic alloys [19-25]. Second, bcc-Fe alloys with Al and/or Cr additions are susceptible to brittle facture [26, 27].

In addition, fuel-cladding tubes in LWRs typically are ~4 m in length with a ~10 mm outside diameter [28]. To

partially compensate for the higher neutron absorption cross section of FeCrAl alloys compared with zirconium

3
alloys, a wall thickness of less than ~0.5 mm is required for FeCrAl tubes. The intrinsic properties of FeCrAl

alloys and the geometry of the fuel cladding in LWRs impose significant challenges for FeCrAl alloy tube

fabrication (e.g., tube drawing and pilgering).

Since the deformability and various physical properties (e.g., yield strength and fatigue and stress corrosion

cracking resistance) of FeCrAl alloys are closely related to their microstructures, microstructural control in the

initial state and during processing is critical for their designed application. The addition of Nb as a minor element

is known to promote the precipitation of a Fe2Nb-type Laves phase in the bcc-Fe matrix [2, 29, 30]. Secondary

particles can have a strong impact on the microstructure evolution (e.g., grain structure and texture) in materials

[31-35]. For example, aluminum nitride particles in aluminum-killed steels retard recrystallization and produce

coarse pancake-shape grain structure with a strong recrystallization texture [32]. However, very few studies [2,

11] have reported investigating how adding Nb to FeCrAl alloys affects their microstructures.

This study examines the microstructures and mechanical properties of deformed and annealed Nb-containing

FeCrAl alloys. The results deepen our understanding of the effects of secondary particles on microstructures and

mechanical properties during the development of advanced alloys. Further guidelines are also provided for

developing FeCrAl alloys to achieve desirable microstructures with balanced processability and physical

properties.

2. EXPERIMENTAL

2.1 Materials Preparation

Four FeCrAl alloys with Nb contents in the range of 0-2 wt% were investigated. Their analyzed compositions are

listed in Table 1. The alloy “2Mo” containing ~2 wt% of Mo instead of Nb was also prepared for comparison. The

alloys were arc-melted by using a pure element feedstock combined with Fe-Y pre-melted alloys (due to the small

Y content), and then drop-cast to make lab-scale bar ingots sized at ~13  25  150 mm. All cast ingots were

homogenized for 1 h at 1200 °C. The 0.7Nb ingot was hot-forged at 1200 °C with a ~60% thickness reduction,

while the 1Nb and 2Nb ingots were hot-rolled at 1000 °C with a ~80% thickness reduction. Then, the Nb-

containing ingots were annealed at 1200 °C for 10 min (0.7Nb) or 1 h (1Nb and 2Nb). The 2Mo ingot was hot-

forged/rolled at 800 °C with a ~70% thickness reduction and then annealed at 800 °C for 0.5 h. These

4
thermomechanical treatments were targeted to form a uniform, equiaxed grain structure with an average grain size

of the magnitude of ~100 μm to eliminate the effect of initial grain size on microstructure evolution during

subsequent process steps. The annealed plates were warm-rolled at 300 °C with a ~10% thickness reduction per

pass, achieving a total ~80% thickness reduction. The final plate thicknesses were ~2.0 mm for 0.7Nb, 0.5 mm for

1Nb and 2Nb, and 1.5 mm for 2Mo. Small test coupons were sectioned from the warm-rolled plates using

electrical discharge machining (EDM) and then isothermally annealed in a temperature range of 600-1100 °C in

air for various periods of time up to 408 h.

2.2 Microstructural characterization

Metallographic samples were mounted in epoxy resin. After grinding, these samples were polished using a

Buehler’s VibroMet polisher with 0.3 μm Al2O3 and subsequent 0.1 μm colloidal silica. Combined

characterization techniques, including light optical microscopy, scanning electron microscopy (SEM), electron

backscatter diffraction (EBSD), and scanning transmission electron microscopy (STEM), were employed. All

micrographs were taken from the transverse direction with the rolling direction (RD) parallel to the horizontal

axis. Etching by a solution of H2O:HNO3:HF = 50:5:1 was applied before optical inspection. A Hitachi S4800

SEM equipped with a back-scattered electron (BSE) detector was used to characterize the grain/subgrain

morphology and Laves phase. EBSD data were collected using a JEOL JSM-6500F SEM and analyzed using

EDAX’s OIM data analysis software package. The scanning step varied from 0.25 to 3 μm. No obvious

microstructural changes were observed across scanned areas. The EBSD data was cleaned up using the “grain

dilation” procedure with a grain tolerance angel of 5° and a minimum grain size of 2-10 μm. Colors in all inverse pole

figure (IPF) maps represented corresponding crystallographic orientations relative to the normal direction (ND) of

the rolled plates based on the color triangle. High-angle boundaries with misorientations larger than 15° were

superimposed on the IPF maps as black lines unless specified otherwise. The grain orientation {hkl}<uvw> was

specified as {hkl} parallel to the rolling plane and <uvw> parallel to the RD. Recrystallized grains were defined as

areas having an average sizes larger than ~5 μm and misorientations smaller than 0.5° between adjacent measured

positions. Texture results were constructed using the spherical harmonics method [36] with a series rank of 16 and

a Gaussian half width of 5°. An orientation tolerance of 15° was used for estimations of the area fractions of

5
certain orientations and texture fibers. Multiple scans were collected in a specimen for statistical accuracy. Table

2 lists the total scanned areas and number of locations for selected specimens. TEM samples were prepared using

an FEI Quanta 3D 200i focused ion beam. An FEI Talos STEM was used for imaging the recovered

microstructures and determining the chemical compositions of the bcc-Fe matrix and the Laves phase.

2.3 Mechanical tests

The Vickers hardness tests were performed using a Shimadzu HMV-G hardness tester with 0.5 kg weight force

and 10 s dwell time. At least six measurements were performed for each specimen. Room-temperature tensile

tests of selected samples were performed with a nominal strain rate of 10-3 s-1. Sub-size dog-bone shape sheet

tensile specimens with a size of ~0.5 × 2 × 10 mm at the gauge section were machined using EDM with the

tensile axis parallel to the RD.

3. RESULTS

3.1 Recrystallization kinetics

Figure 1 plots the recrystallized area fractions of 0.7Nb, 1Nb, and 2Mo as a function of annealing time in a

temperature range of 650-900 °C. For a given alloy, the recrystallization became slow with decreasing annealing

temperature. The recrystallization kinetics also became sluggish as the Nb content increased up to 1 wt%. For

example, 2Mo showed fully recrystallized microstructure after annealing for 30 min at 800 °C, whereas 1Nb still

exhibited partially recrystallized microstructures (~0.81 recrystallized area) even after annealing for 241 h at

800 °C. On the other hand, no significant difference in recrystallization kinetics was observed between 1Nb and

2Nb based on optical micrographs.

3.2 Deformed state

Figure 2a displays a representative IPF color map of as-rolled specimens. The color orientation triangle for IPF

maps is included on the right in Fig. 2a and is the same color orientation triangle for all the following IPF maps.

The as-rolled microstructure consisted of elongated grains along the RD with a width of ~10-50 μm. Because of

the poor indexing of EBSD patterns in severely deformed areas in the as-rolled state, a 1Nb specimen annealed

for 96 h at 700 °C (~0.05 recrystallized area; EBSD scanning area ~0.86 mm2) was used to identify the

deformation texture. As demonstrated in the 𝜑2 = 45° orientation distribution function (ODF) map in Fig. 3, three

6
main texture fibers were presented: α fiber (i.e., <110>//RD), γ fiber (i.e., <111>//ND, blue in IPF maps), and

<100>//ND fiber (red in IPF maps). This finding agrees with the reported deformation texture in low-carbon and

stainless ferritic steels [37-39]. The partial α fiber was extended from {100}<110> to {111}<110>, while γ and

<100>//ND fibers spread in whole orientations. The area fractions of α, γ, and <100>//ND fibers were ~0.50 ±

0.10, 0.39 ± 0.06, and 0.18 ± 0.07, respectively.

Intra-granular and inter-granular deformation heterogeneity was observed in Fig. 2a. The heavily deformed A area

in Fig. 2a was subdivided into ribbon-like structures with high-angle grain boundaries between neighboring

bands. Generally, grains belonging to α (except {111}<110>) and <100>//ND fibers (e.g., B and C areas in

Fig. 2a) showed a more uniform orientation spread than γ-fiber grains (e.g., D and E areas in Fig. 2a). Shear

bands, aligned at ~20-30° to the RD, marked by arrows, were found in the D and E areas in Fig. 2a. EBSD data

within these shear bands were poorly indexed owing to large localized shearing [40]. The distance between shear

bands in the D area was much coarser than that in the E area. Some shear bands in the D area seemed to extend to

neighboring grains. Regions between shear bands in the D area showed a uniform orientation spread, similar to

that found in the B and C areas.

Laves phase particles with a volume fraction of 1.6 ± 0.2% and an equivalent radius of 1.2 ± 0.4 μm were found

in as-rolled 2Nb, as shown in Fig. 2b. However, there was only a small amount of Laves phase in as-rolled 1Nb.

Since warm-rolling at 300 °C would not cause the formation of such Laves phase particles, they would have

existed before the rolling. These facts suggest that the Nb solution limit in the bcc-Fe matrix at 1200 °C was close

to 1 wt% in the examined FeCrAl alloys. Cracks within Laves phase particles and deboning of surfaces between

the matrix and the Laves phase were occasionally observed, as marked by arrows in Fig. 2b. On the other hand,

the Mo-containing 2Mo having a solid-solution bcc matrix exhibited no such Laves phase but had few yttrium-

enriched particles. The Vickers hardness values of as-rolled samples increased with increasing Nb content: 382 ±

11 (0.7Nb), 394 ± 6 (1Nb), 423 ± 8 (2Nb), and 359 ± 12 (2Mo).

3.3 Recovery and recrystallization

Representative microstructural characteristics of 1Nb annealed at 700 °C are summarized in Fig. 4. After

annealing at 700 °C for 1 h, recovery was mostly completed in heavily deformed areas, and most EBSD patterns

7
in such areas were well indexed. One recovered area belonging to the γ fiber is shown in Fig. 4a as an example.

Elongated dislocation cell bands ~0.5 μm in width were aligned at ~35° to the RD, as indicated by the red arrow

line in Fig. 4a. Several distinct bands a few microns wide (S1-S5), intersecting with cell bands, mostly consisted

of well-defined subgrains enclosed by boundaries with misorientations larger than 5°. These bands (S1-S5) likely

had evolved from shear bands. These microstructural features are similar to those in cold-rolled aluminum alloys

[41]. The angles between the S1-S5 bands and the RD were ~35° for S1 and S2 and 10-20° for S3-S5. The

materials in the S1-S5 bands predominantly rotated to around the Goss component (i.e., {110}<100>) and γ fiber.

These bands could act as preferred sites for recrystallization nucleation. It is interesting to notice that shear bands

in γ-fiber grains have often been found to give rise to the Goss and γ-fiber nuclei in ferrous alloys [40, 42, 43],

which is in agreement with the dominant subgrain orientations in the S1-S5 bands. At the same time, a few areas

other than S1-S5 bands also rotated heavily relative to their surroundings.

Upon further annealing for 96 h at 700 °C (Fig. 4b), mixed microstructures consisting of recovered and

recrystallized grains formed. Most recrystallized grains, indicated by arrows in Fig. 4b, nucleated at heavily

deformed regions in γ-fiber grains. Recovered grains belonging to α (except {111}<110>) and <100>//ND fibers,

such as the A and B areas in Fig. 4b, predominantly showed a uniform and limited amount of orientation spread

throughout the grains. The point-to-point misorientation profiles along two lines (L1 and L2) in the A and B areas

are shown in Fig. 4c. As can be seen from Fig. 4c, the misorientation distributions along L1 and L2 were mostly

within 3-4°, together with several peaks of no more than 10°. The average misorientations along L1 and L2 were

~1.5° and 1.0°, respectively. On the other hand, well-defined “polygonised grains” were predominantly observed

in γ-fiber recovered grains. Figure 4d qualitatively describes the boundary characteristics of these polygonised

grains. As shown in Fig. 4d, most of these grains (~1-2 μm in size, smaller than the threshold value of

recrystallized grains) were disoriented relative to one another by misorientations larger than 15°; whereas the

misorientation within these grains was mostly smaller than 0.5°, similar to that within the recrystallized grains on

the right side. These findings indicate that these well-defined polygonised grains in γ-fiber grains experienced

extensive recovery. Recrystallized grains were often observed around polygonised grains. As is discussed in

Section 4.2, Laves phase particles in 1Nb strongly impeded the nucleation and growth of recrystallized grains.

8
However, some special boundaries may remain unpinned and mobile owing to their lower energy compared with

other high-angle boundaries, as suggested by Somerday and Humphreys [35] in their study of the recrystallization

behavior of supersaturated Al-Mn alloys. Hence, some polygonised grains with special boundaries relative to their

surroundings were likely to serve as recrystallization nucleation sites.

The incubation time for recrystallization largely decreased with increasing annealing temperature, as indicated by

Fig. 1. Representative microstructures of 1Nb subjected to isothermal annealing in the range of 800-1100 °C are

shown in Fig. 5. The annealing at 800 °C for 24 h and at 900 °C for 1 h (Figs. 5a and 5b) produced a similar

recrystallized area fraction of ~0.75. Some areas seemed to have a greater tendency to remain non-recrystallized

than others. Non-recrystallized areas, marked by arrows in Figs 5a and 5b, usually featured a uniform and small

misorientation spread. At both temperatures, recrystallized grains were preferentially elongated along the RD,

which seemed to be confined by pre-existing grain boundaries. Mechanisms to form the elongated recrystallized

grains are correlated with the Laves phase precipitations, and the details will be further discussed in Section 4.2.

The average recrystallized grain sizes were 51 ± 30 μm (24 h at 800 °C) and 26 ± 14 μm (1 h at 900 °C) with

grain aspect ratios in a range from 0.28 to 0.35. The phenomenon is contrary to the classical view regarding the

relationship between recrystallized grain sizes and annealing temperatures; possible mechanisms will be discussed

later. Annealing at temperatures above 1100 °C led to coarse, equiaxed grains in 1Nb. One example is shown in

Fig. 5c. After annealing for 0.5 h at 1100 °C, the average grain size was 82 ± 36 μm with a grain aspect ratio of

0.56 ± 0.11.

Figure 6 compares the representative recrystallization textures of 2Mo and 1Nb. As shown in Fig. 6a, 2Mo

displayed a strong γ fiber texture with an area fraction of ~0.43 after annealing at 800 °C, which is in agreement

with the recrystallization texture in low-carbon and ferritic stainless ferritic steels [32, 38, 39]. On the other hand,

as shown in Figs 6b and 6c, recrystallized grains in 1Nb annealed at 800 and 900 °C showed weak texture. The

area fractions of recrystallized γ fiber were ~0.11 (24 h/800 °C) and 0.19 (1 h/900 °C). However, the annealing at

1100 °C produced stronger γ fiber in 1Nb with an area fraction of ~0.35 after annealing for 0.5 h at 1100 °C.

Possible mechanisms governing the recrystallization texture formation will be discussed later.

3.4 Laves phase particles

9
In the high-angle annular dark-field (HAADF) image in Fig. 7a, subgrains with a size of ~0.5 μm were found in

1Nb after annealing for 8 h at 700 °C. Boundaries in Fig. 7a were decorated by particles with irregular shapes

(bright contrast). The compositions of the matrix and particles were measured and listed in Table 3. As shown in

Table 3, the Fe/Nb/Cr/Al/Si atomic ratio was ~1.82:1:0.24:0.12:0.25, close to that of the Fe2Nb compound,

confirming the formation of a Fe2Nb-type Laves phase during annealing. Besides, the EBSD patterns from those

particles were indexed, showing the Fe2Nb-type structure. Laves phase particles contained less Cr and Al but were

enriched in Si compared with the matrix. The distribution of Laves phase particles depended on the matrix

characteristics, as shown in Figs 8a and b. Large particles (> ~100 nm) were predominantly located along

subgrain boundaries in the upper half of Fig, 8a. By contrast, the microstructure in the lower half of Fig. 8a had

ill-defined subgrains decorated with high-density, fine-scale Laves phase particles. It appears that advancing

subgrain boundaries can sweep up fine particles, leaving coarser ones on the migrating boundaries [33]. Laves

phase particles in 1Nb formed with similar distributions in both non-recrystallized and recrystallized areas in the

early stage of recrystallization at and below 900 °C, suggesting that the nucleation of Laves phase particles

occurred before recrystallization at these temperatures. As the recrystallization process proceeded, most particles

migrated onto grain boundaries, as shown in Fig. 8b. It should be emphasized that some non-recrystallized regions

consisting of subgrains (~2-5 μm in size) still existed after annealing for 72 h at 900 °C.

Figure 9 plots the volume fraction of the Laves phase in 0.7Nb and 1Nb as a function of annealing temperature. In

Fig. 9, specimens annealed for sufficient long periods at corresponding temperatures were analyzed, and the

measured particle volume fractions should be close to the equilibrated ones based on the observed results. In both

alloys, the volume fraction of the Laves phase decreased as the annealing temperature increased in the studied

range. For example, the volume fraction of the Laves phase in 1Nb decreased from 2.54 ± 0.22% at 700 °C to

0.35 ± 0.04% at 1100 °C. Meanwhile, 0.7Nb had a smaller amount of Laves phase than 1Nb for a given

annealing temperature. For instance, at 900 °C, the particle volume fraction of 1Nb was 1.82 ± 0.09%, whereas it

was only 1.19 ± 0.22% in 0.7Nb.

3.5 Mechanical properties of deformed and annealed samples

10
Figure 10a displays the annealing effect on the Vickers hardness of 1Nb in the range of 600-900 °C. Overall, the

Vickers hardness value dropped as the annealing time increased, and it was lower at higher temperatures for a

given annealing time. The curves of 600, 650, and 700 °C had plateaus at which the hardness value slightly

decreased or stayed constant. The plateau starting points shifted to longer times with decreasing annealing

temperatures. The residual strength of recovered metals has been reported to logarithmically decay as a function

of annealing time [44]. Two factors probably contribute to these plateau formations. First, Laves phase particles

slow down the recovery process by impeding thermally activated dislocation movements (e.g., dislocation

gliding). Next, the Laves phase strengthens the recovered FeCrAl alloys through particle hardening. Figure 10b

compares the annealing effect on the Vickers hardness of 0.7Nb, 1Nb, 2Nb, and 2Mo at 700 °C. As shown in

Fig. 10b, the hardness of 2Mo quickly decreased upon annealing at 700 °C compared with the hardness in the as-

rolled condition; and it had the lowest hardness for a given annealing time among the studied alloys, which agrees

with the recrystallization kinetics shown in Fig. 1. For Nb-containing FeCrAl alloys, the Vickers hardness value

increased with the increasing Nb content for a given annealing time.

Figure 10c shows tensile engineering stress vs. strain curves of selected 1Nb samples. The as-rolled 1Nb

exhibited the highest yield stress of ~1300 MPa; however, the stress almost immediately decreased after yielding

with little uniform strain. The yield stress values of annealed specimens decreased, corresponding to the hardness

drop in Fig. 10a. Meanwhile, the deformability of 1Nb was improved. For example, 1Nb exhibited a yield stress

of ~540 MPa with a uniform strain of ~10.4% after annealing for 24 h at 900 °C. The alloy 0.7Nb annealed at 950

°C for 0.5 h was reported to show similar tensile properties at room temperature with a yield stress ~500 MPa and

a uniform strain of ~16.2% [11]. In comparison, the yield stress values of 2Mo were in the range from ~440 to

480 MPa as the average grain size decreased from ~165 to 30 μm [11]. The strength of base FeCrAl alloys with

10-20 wt% Cr and 3-5 wt% Al further dropped with the yield stress values in the range of ~350-450 MPa [2].

Meanwhile, the deformability of FeCrAl alloys showed strong correlation with their microstructures. For

example, sudden, brittle fracture with small plastic elongation occurred in room-temperature tensile tests of 2Mo

when the average grain size was larger than ~80 μm [11], suggesting the importance of microstructural control to

achieve the desirable mechanical properties of FeCrAl alloys.

11
4. Discussion

4.1 Orientation-dependent microstructures in deformed and annealed materials

As demonstrated in Figs. 2, 4, and 5, the microstructural evolution in deformed and annealed Nb-containing

FeCrAl alloys showed a strong dependence on crystallographic orientations. The orientation dependence can be

described on the basis of the Taylor factor, M, as defined by [45-48]:


𝜎𝑖𝑗 ∑ 𝛿𝛾𝑘
𝑀= = , (1)
𝜏𝑐 𝛿𝜀𝑖𝑗

where 𝛿𝜀𝑖𝑗 represents a plastic strain increment under a state of stress 𝜎𝑖𝑗 , 𝜏𝑐 is the critical resolved shear stress,

and ∑ 𝛿𝛾𝑘 is the sum of shear strains on given slip systems to accomplish the strain 𝛿𝜀𝑖𝑗 . The M value map for the

rolling deformation with φ2 = 45° is displayed in Fig. 11. As shown in Fig. 11, <100>//ND-fiber grains have small

M values of ~2.45, whereas γ-fiber grains possess larger M values from 3.67 ({111}<112>) to 4.08

({111}<110>). The M values of grains belonging to α fiber increase from 2.45 to 4.08 as the orientation shifts

from {100}<110> to {111}<110>. Based on Eq. (1), grains with small M values are expected to more easily

accommodate applied strain (smaller ∑ 𝛿𝛾𝑘 for a given 𝛿𝜀𝑖𝑗 ) than grains with large M values.

The dependence of deformed microstructures on grain orientations in low-carbon steels and pure iron is reported

by numerous researchers. Dillamore et al. [49] studied the deformed microstructures in low-carbon steels and pure

iron with a 70% thickness reduction. They demonstrated that the subgrain size decreased from ~1 μm to 0.4 μm

and its misorientation increased from ~1° to 6° as the plane shifted from {100} to {110} along the α fiber, while

γ-fiber grains other than those oriented around {111}<110> had a similar subgrain size and misorientation to

{110}<110>-oriented grains. Samajdar et al. [50] and Haldar et al. [51] concluded similarly for deformed

interstitial-free steels. The elastic stored energy, V, was also found to be orientation-dependent (relative to the

ND), and the sequence was V110 > V111 > V211 > V100 [49, 50, 52]. Furthermore, it was reported that intragranular

shear bands in low-carbon steels preferentially occurred in grains with large M values (e.g., γ-fiber grains and

{110}<110>-oriented grains) [37, 43].

In the study, most of the α-fiber grains ranging from {100}<110> to {112}<110> and <100>//ND grains showed

relatively uniform deformation throughout the grains (Fig. 4b). By contrast, the deformation in γ-fiber grains was

12
highly heterogeneous. Heavily deformed shear bands were dominantly observed in γ-fiber grains, while areas

other than the shear bands in γ-fiber grains tended to experience much less deformation with a smooth orientation

spread (Figs. 2a and 4a).

As shown in Fig. 4b, recrystallized grains in 1Nb were mainly found in γ-fiber grains at an early stage of

recrystallization owing to their high stored energy and strong deformation inhomogeneity [50, 52, 53]. At a higher

annealing temperature (e.g., 900 °C), the advantage of preferred nucleation in γ-fiber grains is likely to become

weaker, since most areas tended to recrystallize in a short period (Fig. 5b). Bands with coarser grains than their

surroundings were usually observed in annealed Nb-containing FeCrAl alloys [11], indicating that some deformed

grains with certain orientations intrinsically had fewer recrystallization nucleation sites. To understand the

recrystallization resistance of grains in different orientations, the main deformation texture is categorized into

three groups with 15° tolerance: <100>//ND fiber, γ fiber, and {112}<110> component. Notice that there is some

overlap between the γ fiber and {112}<110> component. Figure 12 plots the area fractions of non-recrystallized

<100>//ND fiber, γ fiber, and the {112}<110> component after annealing at various conditions. Overall, as the

recrystallization process proceeded, the area fraction of γ fiber and {112}<110> component increased, while the

area fraction of <100>//ND fiber decreased. Based on the results, {112}<110> component had the strongest

tendency to remain non-recrystallized among the analyzed texture components in 1Nb. Meanwhile, not all γ-fiber

areas showed a similar tendency to recrystallize; as a result of the strong deformation inhomogeneity in γ fiber,

some γ-fiber areas (probably ones between shear bands) remained non-recrystallized even after extended

annealing. It is worth noting that the formation of polygonised grains in 1Nb at 700 °C (Figs. 4b and 4d) largely

decreased the stored energy of heavily deformed areas in γ-fiber grains. The <100>//ND-fiber grains could have

experienced higher deformation than other grains because of their small M value (~ 2.45). Samajdar et al. [50]

also reported that {100}<110>-oriented grains show a larger spread of stored energies than {112}<110>-oriented

ones. Both findings may explain the area fraction decrease of <100>//ND-fiber grains as the recrystallization

process proceeded, as indicated in Fig. 12.

4.2 Recrystallized grain size vs. annealing temperature

13
As stated in Section 3.3, 1Nb exhibited an abnormal relationship between recrystallized grain size and annealing

temperature. Figure 13 plots the recrystallized grain size in 2Mo and 1Nb after annealing at various conditions.

The 2Mo showed fully recrystallized grain structure after annealing for 0.5 h at 750 °C. As shown in Fig. 13, the

recrystallized grain size in 2Mo increased as annealing temperature increased after 0.5 h annealing, which agrees

with the classical relationship between recrystallized grain size and annealing temperature. For 1Nb, on the other

hand, the recrystallized grain size was stable at ~25 μm at 900 °C even after annealing for 24 h; whereas the same

alloy had a recrystallized grain size (~20-23 μm) at a lower annealing temperature (i.e., 700 and 800 °C) in the

early stage of recrystallization, similar to that at 900 °C. However, recrystallized grains coarsened as annealing

time increased. For instance, the recrystallized grain size increased to ~38 μm in 1Nb after annealing for 408 h at

700 °C (recrystallized area fraction ~0.26). Furthermore, recrystallized grains in 1Nb could continue to grow by

advancing into nearby non-recrystallized areas at 700 and 800 °C, leading to even coarser grains than those shown in

Fig. 13. Therefore, it is concluded that annealing at 900 °C is optimal for maintaining a fine grain structure in the

alloy 1Nb, compared with the structure at lower temperatures, given a similar recrystallized area fraction.

Two factors govern the recrystallized grain size: recrystallization nucleation and grain growth rates. Both are

strongly influenced by the pinning effect of Laves phase particles on subgrain and grain boundaries in Nb-

containing FeCrAl alloys. Assuming that particles mainly lie on subgrain/grain boundaries, the pinning pressure

provided by Laves phase particles, Pz, can be expressed as [54]

3𝛾𝑓𝑙
𝑃𝑧 = 2𝜋𝑟2 , (2)

where γ is the interfacial energy per unit area of boundary, f is the particle volume fraction, r is the average

particle radius, and l is the average grain intercept length. The average grain intercept length can be related to the

average grain diameter, D, by 𝑙 = 𝜋𝐷/4 [55].

The driving pressure of the recrystallization-front migration, Pr, can be related to the dislocation density

difference (∆𝜌) between recrystallized and non-recrystallized regions by [54, 56]

𝑃𝑟 = 𝐺𝑏 2 ∆𝜌/2 , (3)

14
where G is the shear modulus (~50 GPa at 800 °C [57]), b is the Burgers vector length (~0.25 nm). The value of

∆𝜌 is estimated to be ~6×1014 /m2 based on the strain hardening model [58]. In the model, assuming that the

average Taylor factor is 3, we have

∆𝜎 ~ 3/2𝐺𝑏√Δ𝜌 , (4)

where ∆𝜎 (~760 MPa in 1Nb, as shown in Fig. 10c) is the yield stress increment between as-rolled and fully

annealed materials, and G ~80 GPa at room temperature for FeCrAl alloys. Substituting for all parameters in Eq.(

4), Pr is estimated to be ~0.9 MPa in 1Nb at 800 °C. For 1Nb after annealing for 1 h at 800 °C, the Laves phase

particle size was ~100 nm with f ~0.02 (Fig. 9) and l of the magnitude of 0.5 μm. The value of γ was reported to

be ~1 J/m2 [59]. Substituting for all parameters in Eq. (2), Pz is determined to be ~1.9 MPa, which is

approximately twice the value of Pr. Consequently, Laves phase particles were able to effectively retard

recrystallization nucleation at 800 °C, leading to a small recrystallization area fraction (~0.17, Fig. 1) in 1Nb after

annealing for 1 h at 800 °C. Furthermore, the extended recovery at 700 and 800 °C in non-recrystallized regions

released the stored elastic energy and thus decreased the driving force for recrystallization. These facts contribute

to the low nucleation rate in 1Nb below 900 °C. The pinning effect of Laves phase particles is expected to be

smaller at higher temperatures (e.g., 900 °C) because of the smaller amount of Laves phase and higher particle

coarsening rate. For example, the value of Pz is ~0.8 MPa for 1Nb annealed for 1 h at 900 °C (f ~0.018 and

particle size ~150 nm). Therefore, more deformed grains were able to overcome the pinning effect of Laves phase

particles and nucleate in 1Nb when it was annealed at 900 °C, as demonstrated in Figs 1 and 5b. As stated in

Sections 3.3 and 4.1, some areas with small misorientation spreads remained non-recrystallized after extended

annealing (Figs. 5 and 12). These areas were likely to have small Pr; thus, the advance of recrystallized grain

boundaries into these areas was largely inhibited.

Regarding recrystallized grain growth, its driving pressure, Pg, can be described as [56]

𝑃𝑔 = 2𝛾/𝐷 . (5)

In this case, assuming that Pz and Pg achieve balance at an average grain size of 25 μm at 900 °C, the threshold

value of r should be ~1.4 μm, based on Eqs. (2) and (5). However, the average equivalent radius of Laves phase

15
particles along grain boundary in 1Nb after annealing for 72 h at 900 °C was ~0.5 μm, indicating that the particles

still acted effectively to resist grain growth. This is why the sizes of recrystallized grains in 1Nb were stable at

900 °C in the studied annealing periods, as shown in Fig. 13. Therefore, the combined effects of Laves phase

particles on the recrystallization nucleation and grain growth rates led to a finer recrystallized grain structure at

900 °C than at lower annealing temperatures, as shown in Fig. 13. A grain structure with an average grain size of

~20 μm was also achieved in 0.7Nb after annealing at 950 °C for 0.5 h [11]. In addition, grain/subgrain

boundaries in rolled materials tended to align with the RD, leading to a preferred distribution of the Laves phase

along the RD. This finding explains the elongated grain morphology seen in Figs. 5a and 5b. The confinement of

most recrystallized grains by pre-existing grain boundaries in Figs. 5a and 5b was probably due to higher particle

density along pre-existing boundaries. At even higher annealing temperatures (e.g., 1100 °C), a coarse equiaxed

grain structure formed (Fig. 5c) because the pinning effect of secondary particles was limited or nonexistent.

4.3 Effect of Laves phase particles on recrystallization texture

The most noticeable feature in the recrystallization texture of Nb-containing FeCrAl alloys compared with 2Mo

was the disappearance of the γ fiber upon annealing at below 1100 °C, as shown in Fig. 6a-c. The texture in

FeCrAl alloys affects not only their processability in thin-wall tube fabrication but also their in-service

performance (e.g., yield strength, thermal and in-reactor creep strength, and fatigue and stress corrosion cracking

resistance) [12, 60]. The recrystallization texture in low-carbon steels has attracted extensive attention because of

the direct relevance of deep drawability to the texture state. Strong γ fiber was usually observed in recrystallized

low-carbon steels. The formation of strong recrystallization γ fiber is often attributed to preferred nucleation

and/or grain growth [43, 50, 52, 61, 62]. In-grain shear bands in γ-fiber deformed grains were demonstrated to

serve as important nucleation sites for γ-fiber recrystallized grains [43, 62]. Emren et al. [61] argued that

{111}<112>-oriented nuclei could favorably grow into nearby {112}<110>-oriented deformed grains because

they are related by a 35° rotation around the <110> axis, which is close to the maximum growth rate of new

grains into a deformed matrix in bcc metals. When annealed below 1100 °C, most recrystallized grains in 1Nb

were restrained by pre-existing grain boundaries, which eliminated the possible preferred grain growth in 1Nb. At

the same time, no areas with dominant γ-fiber recrystallized grains were observed in 1Nb annealed below 1100

16
°C, indicating the suppression of the γ-fiber nuclei in 1Nb by Laves phase particles. Even if there were some

preferred γ-fiber nuclei, they could not quickly advance into neighboring areas because of the pinning effect of the

Laves phase on grain boundaries. Therefore, neither nucleation nor growth of special orientation was preferred

with the strong pinning effect of the Laves phase, which led to the disappearance of prevalent recrystallized γ-

fiber grains in 1Nb (Figs. 6b and 6c). This also can explain why a considerable number of γ-fiber grains were

observed in 1Nb annealed at 1100 °C when fewer Laves phase particles were present.

4.4 Further development of Nb-containing FeCrAl alloys

As stated in Section 3.2, the Nb solubility in the bcc-Fe matrix at 1200 °C was ~1 wt% in the studied FeCrAl

alloys. FeCrAl alloys with a Nb content higher than 1 wt% are likely to have micron-sized Laves phase particles

after thermo-mechanical processing, as shown in Fig. 2b. These large Laves phase particles have limited pinning

effects on grains/subgrains and probably lead to micro-cracks in FeCrAl alloys. In view of these facts, a Nb

content below 1 wt% would be preferable in further development of Nb-containing FeCrAl alloys.

As shown in Fig. 9, the volume percentage of Laves phase in 1Nb at 1100 °C was ~0.35 ± 0.04%, and Laves

phase particles had a limited effect on the stabilization of grains at and above 1100 °C (Fig. 5c). Meanwhile, the

pinning effect of the Laves phase decreased as the Nb content decreased from 1 wt% at a given temperature, as

demonstrated by Fig. 1. Therefore, it would be difficult to control the microstructures of FeCrAl alloys above

1100 °C only through Nb addition. On the other hand, for hot or warm working below 1100 °C, the Laves phase

in Nb-containing FeCrAl alloys can reduce the recovery during processing and thus increase the accumulated

stored energy for subsequent annealing treatments. Grain refinement in Nb-containing FeCrAl alloys could be

achieved through properly designed thermomechanical processing, as suggested by Figs 5b and 13, which would

improve the processability as well as the mechanical properties of FeCrAl alloys. At the same time, in the multi-

pass tube fabrication process, Laves phase particles could suppress grain coarsening rates during inter-pass

annealing, helping obtain desirable microstructures for balanced deformability and the final properties of tube

products. In addition, the disappearance of strong recrystallization of γ fiber along the thickness direction in Nb-

containing FeCrAl alloys would benefit thickness reduction during tube fabrication.

17
5. Conclusions

In this study, the microstructures and mechanical properties of as-rolled and annealed Nb-containing FeCrAl

alloys are investigated. Based on the results, the microstructures of FeCrAl alloys can be tailored through Nb

addition to achieve a balance between processability and physical properties for their desired application as fuel

cladding in LWRs. Suggestions for further development of Nb-containing FeCrAl alloys are also provided at the

end. The following are conclusions of the study:

(1) Fe2Nb-type Laves phase particles stabilized the microstructures in Nb-containing FeCrAl alloys because of

their fine dispersion along grain/subgrain boundaries. The deformation microstructures were more stable with

increasing Nb content up to 1 wt%. The pinning effect of the Laves phase decreased with increasing annealing

temperature. Above 1100 °C, the pinning effect of the Laves phase in 1Nb was negligible.

(2) Deformation texture typical in low-carbon and ferritic stainless steels was observed in the studied FeCrAl

alloys. The main texture fibers included α, γ, and <100>//ND fibers. Strong deformation heterogeneity was

observed in γ-fiber grains with heavily deformed shear bands, while <100>//ND fiber-grains and {112}<110>-

oriented grains tended to have uniform deformation throughout the grains.

(3) In 1Nb, recrystallization grains preferentially nucleated at heavily deformed areas with large orientation

spreads, such as shear bands in γ-fiber grains. On the other hand, areas with uniform and small misorientation

spreads, including some γ-fiber areas, tended to remain non-recrystallized even after extended annealing. The

{112}<110>-oriented grains showed the strongest resistance to recrystallizing among the analyzed texture

components in 1Nb.

(4) Recrystallized grains in 1Nb, most of them confined by pre-existing grain boundaries, were elongated along

the RD after annealing below 1100 °C. An abnormal relationship between recrystallized grain size and annealing

temperature was observed in 1Nb. Annealing at 900 °C resulted in a finer grain structure in 1Nb than annealing at

700 and 800 °C for a given recrystallization area fraction.

(5) Recrystallized grains in Nb-containing FeCrAl alloys showed weak texture with the disappearance of

prevalent γ-fiber grains, which was the opposite of the characteristics of the Mo-containing FeCrAl alloy and low-

18
carbon steels. Such weak texture formation would promote relatively easy deformation, which should support

improved processability of FeCrAl alloys.

19
Acknowledgments

The authors wish to thank Cecil Carmichael, Gregory Cox, David Harper, Kevin Hanson, Dustin Heidel, and

Daniel Moore for their help in materials preparation and processing. We also appreciate Kurt Terrani’s support as

the program manager. This research was funded by the US Department of Energy’s Office of Nuclear Energy,

Advanced Fuel Campaign of the Fuel Cycle R&D program.

20
REFERENCES

1. K.A. Terrani, S.J. Zinkle, L.L. Snead, Advanced oxidation-resistant iron-based alloys for LWR fuel cladding, J.

Nucl. Mater. 448 (2014) 420-435.

2. Y. Yamamoto, B.A. Pint, K.A. Terrani, K.G. Field, Y. Yang, L.L. Snead, Development and property evaluation of

nuclear grade wrought FeCrAl fuel cladding for light water reactors, J. Nucl. Mater. 467 (2015) 703-716.

3. S.A. Briggs, P.D. Edmondson, K.C. Littrell, Y. Yamamoto, R.H. Howard, C.R. Daily, K.A. Terrani, K. Sridharan,

K.G. Field, A combined APT and SANS investigation of α′ phase precipitation in neutron-irradiated model FeCrAl

alloys Acta Mater. 129 (2017) 217-228.

4. K.G. Field, S.A. Briggs, K. Sridharan, R.H. Howard, Y. Yamamoto, Mechanical properties of neutron-irradiated

model and commercial FeCrAl alloys, J. Nucl. Mater. 489 (2017) 118-128.

5. K.G. Field, M.N. Gussev, Y. Yamamoto, L.L. Snead, Deformation behavior of laser welds in high temperature

oxidation resistant Fe–Cr–Al alloys for fuel cladding applications, J. Nucl. Mater. 454 (2014) 352-358.

6. J.C. Haley, S.A. Briggs, P.D. Edmondson, K. Sridharan, S.G. Roberts, S. Lozano-Perez, K.G. Field, Dislocation

loop evolution during in-situ ion irradiation of model FeCrAl alloys, Acta Mater. 136 (2017) 390-401.

7. M. Moalem, D.R. Olander, Oxidation of Zircaloy by steam, J. Nucl. Mater. 182 (1991) 170-194.

8. R.E. Pawel, J.V. Cathcart, R.A. McKee, The kinetics of oxidation of Zircaloy-4 in steam at high temperatures, J.

Electrochem. Soc. 126 (1979) 1105-1111.

9. B.A. Pint, K.A. Terrani, M.P. Brady, T. Cheng, J.R. Keiser, High temperature oxidation of fuel cladding candidate

materials in steam–hydrogen environments, J. Nucl. Mater. 440 (2013) 420-427.

10. Z.G. Zhang, F. Gesmundo, P.Y. Hou, Y. Niu, Criteria for the formation of protective Al 2O3 scales on Fe–Al and Fe–

Cr–Al alloys, Corros. Sci. 48 (2006) 741-765.

11. Z. Sun, H. Bei, Y. Yamamoto, Microstructural control of FeCrAl alloys using Mo and Nb additions, Mater. Charact.

132 (2017) 126-131.

12. Z. Sun, Y. Yamamoto, Processability evaluation of a Mo-containing FeCrAl alloy for seamless thin-wall tube

fabrication, Mater. Sci. Eng. A 700 (2017) 554-561.

13. Y. Yamamoto, Z. Sun, Quality optimization of commercial FeCrAl tube production. 2017, Oak Ridge National

Laboratory, M3FT-17OR020202121, ORNL/TM-2017/338.

21
14. Y. Yamamoto, Z. Sun, M.N. Gussev, B.A. Pint, K.A. Terrani, Examination of compressive deformation routes for

production of ATF FeCrAl tubes. 2016, Oak Ridge National Laboratory, M3FT-16OR020202133, ORNL/TM-

2016/509.

15. Y. Yamamoto, Z. Sun, B.A. Pint, K.A. Terrani, Optimized Gen-II FeCrAl cladding production in large quantity for

campaign testing. 2016, Oak Ridge National Laboratory, M3FT-16OR020202132, ORNL/TM-2016/227.

16. O. Kubaschewski, Iron-binary Phase Diagrams. 1982: Springer-Verlag Berlin Heidelberg.

17. R. Rana, C. Liu, R.K. Ray, Low-density low-carbon Fe–Al ferritic steels, Scr. Mater. 68 (2013) 354-359.

18. D.-W. Suh, N.J. Kim, Low-density steels, Scr. Mater. 68 (2013) 337-338.

19. G. Song, Z. Sun, B. Clausen, P.K. Liaw, Microstructural characteristics of a Ni 2TiAl-precipitate-strengthened

ferritic alloy, J. Alloys Compd. 693 (2017) 921-928.

20. G. Song, Z. Sun, L. Li, B. Clausen, S.Y. Zhang, Y. Gao, P.K. Liaw, High Temperature Deformation Mechanism in

Hierarchical and Single Precipitate Strengthened Ferritic Alloys by In Situ Neutron Diffraction Studies, Sci. Rep. 7

(2017) 45965.

21. G. Song, Z. Sun, L. Li, X. Xu, M. Rawlings, C.H. Liebscher, B. Clausen, J. Poplawsky, D.N. Leonard, S. Huang, Z.

Teng, C.T. Liu, M.D. Asta, Y. Gao, D.C. Dunand, G. Ghosh, M. Chen, M.E. Fine, P.K. Liaw, Ferritic alloys with

extreme creep resistance via coherent hierarchical precipitates, Sci. Rep. 5 (2015) 16327.

22. G. Song, Z. Sun, J.D. Poplawsky, Y. Gao, P.K. Liaw, Microstructural evolution of single Ni 2TiAl or hierarchical

NiAl/Ni2TiAl precipitates in Fe-Ni-Al-Cr-Ti ferritic alloys during thermal treatment for elevated-temperature

applications, Acta Mater. 127 (2017) 1-16.

23. Z. Sun, C.H. Liebscher, S. Huang, Z. Teng, G. Song, G. Wang, M. Asta, M. Rawlings, M.E. Fine, P.K. Liaw, New

design aspects of creep-resistant NiAl-strengthened ferritic alloys, Scr. Mater. 68 (2013) 384-388.

24. Z. Sun, G. Song, J. Ilavsky, G. Ghosh, P.K. Liaw, Nano-sized precipitate stability and its controlling factors in a

NiAl-strengthened ferritic alloy, Sci. Rep. 5 (2015) 16081.

25. Z. Sun, G. Song, J. Ilavsky, P.K. Liaw, Duplex pecipitates and their effects on the room-temperature fracture

behaviour of a NiAl-strengthened ferritic alloy, Mater. Res. Lett. 3 (2015) 128-134.

26. J. Herrmann, G. Inden, G. Sauthoff, Deformation behaviour of iron-rich iron-aluminum alloys at low temperatures,

Acta Mater. 51 (2003) 2847-2857.

22
27. H.P. Qu, Y.P. Lang, C.F. Yao, H.T. Chen, C.Q. Yang, The effect of heat treatment on recrystallized microstructure,

precipitation and ductility of hot-rolled Fe–Cr–Al–REM ferritic stainless steel sheets, Mater. Sci. Eng. A 562 (2013)

9-16.

28. S.J. Zinkle, G.S. Was, Materials challenges in nuclear energy, Acta Mater. 61 (2013) 735-758.

29. E. Hornbogen, Precipitation from binary substitutional solid solutions of alpha iron, in Precipitation from Iron-base

Alloys, G.R. Speich and J.B. Clark, Editors. 1965, Gordon and Breach Science Publishers: NY. p. 1-67.

30. G.R. Speich, Precipitation of Laves phases from iron-niobium (columbium) and iron-titanium solid solution, Trans.

Metall. Soc. AIME 224 (1962) 850-858.

31. A.S. Keh, W.C. Leslie, G.R. Speich. Precipitation on substructure in iron-base alloys. in Symposium on the role of

substructure in the mechanical behavior of metals. 1962.

32. W.B. Hutchinson, Development and control of annealing textures in low-carbon steels, Int. Mater. Rev. 29 (1984)

25-42.

33. W.B. Hutchinson, B.J. Duggan, Influence of precipitation on recrystallization and texture development in an iron-

1.2% copper alloy, Met. Sci. 12 (1978) 372-380.

34. F.J. Humphreys, A unified theory of recovery, recrystallization and grain growth, based on the stability and growth

of cellular microstructures—II. The effect of second-phase particles, Acta Mater. 1997 (1997) 5031-5039.

35. M. Somerday, F.J. Humphreys, Recrystallisation behaviour of supersaturated Al–Mn alloys Part 1 – Al–1.3 wt-

%Mn, Mater. Sci. Tech. 19 (2014) 20-29.

36. H.J. Bunge, Texture Analysis in Materials Science: Mathematical Methods. 1982, London: Butterworths.

37. B. Hutchinson, Deformation microstructures and textures in steels, Phil. Trans. R. Soc. Lond. A 357 (1999) 1471-

1485.

38. D. Raabe, K. Lücke, Textures of ferritic stainless steels, Mater. Sci. Tech. 9 (1993) 302-312.

39. D. Raabe, K. Lücke, Rolling and annealing textures of bcc metals, Mater. Sci. Forum 157 (1994) 597-610.

40. K. Ushioda, W.B. Hutchinson, Role of shear bands in annealing texture formation in 3%Si–Fe(111)[112] single

crystals, ISIJ Int. 29 (1989) 862-867.

41. P.J. Hurley, F.J. Humphreys, The application of EBSD to the study of substructural development in a cold rolled

single-phase aluminium alloy, Acta Materialia 51 (2003) 1087-1102.

23
42. T. Haratani, W.B. Hutchinson, I.L. Dillamore, P. Bate, Contribution of shear banding to origin of Goss texture in

silicon iron, Met. Sci. 18 (2013) 57-66.

43. M.R. Barnett, Role of in-grain shear bands in the nucleation of <111>//ND recrystallization textures in warm rolled

steel, ISIJ Int. 38 (1998) 78-85.

44. E. Nes, Recovery revisited, Acta Metall. Mater. 43 (1995) 2189-2207.

45. C.N. Reid, Deformation Geometry for Materials Scientists. International Series on Materials Science and

Technology. Vol. 11. 1973, New York: Pergamon Press.

46. J.F.W. Bishop, R. Hill, XLVI. A theory of the plastic distortion of a polycrystalline aggregate under combined

stresses, Phil. Mag. 42 (1951) 414-427.

47. J.F.W. Bishop, R. Hill, CXXVIII. A theoretical derivation of the plastic properties of a polycrystalline face-centred

metal, Phil. Mag. 42 (1951) 1298-1308.

48. G.I. Taylor, Plastic strain in metals, J. Inst. Metals 62 (1938) 307-324.

49. I.L. Dillamore, C.J.E. Smith, T.W. Watson, Oriented Nucleation in the Formation of Annealing Textures in Iron,

Metal Science Journal 1 (1967) 49-54.

50. I. Samajdar, B. verlinden, P. Van Houtte, D. Vanderschueren, γ-Fibre recrystallization texture in IF-steel: an

investigation on the recrystallization mechanisms, Mater. Sci. Eng. A 238 (1997) 343-350.

51. A. Haldar, X. Huang, T. Leffers, N. Hansen, R.K. Ray, Grain orientation dependence of microstructures in a warm

rolled IF steel, Acta Materialia 52 (2004) 5405-5418.

52. R.L. Every, M. Hatherly, Oriented nucleation in low-carbon steels, Texture 1 (1974) 183-194.

53. D. Raabe, Investigation of the orientation dependence of recovery in low-carbon steel by use of single orientation

determination, Steel Res. 66 (1995) 222-229.

54. S.S. Hansen, J.B. Vander Sande, M. Cohen, Niobium carbonitride precipitation and austenite recrystallization in

hot-rolled microalloyed steels, Metall. Trans. A 11 (1980) 387-402.

55. ASTM, Standard test methods for determining average grain size, E112-12. 2012, ASTM International: West

Conshohocken, PA.

56. P.A. Manohar, M. Ferry, T. Chandra, Five decadesof the zener equation, ISIJ Int. 38 (1998) 913-924.

57. M. Fukuhara, A. Sanpei, Elastic moduli and internal friction of low carbon and stainless steels as a function of

temperature, ISIJ Int. 33 (1993) 508-512.

24
58. G.E. Dieter, Mechanical Metallurgy. Vol. 3. 1986, New York: McGraw-hill.

59. S. Ratanaphan, D.L. Olmsted, V.V. Bulatov, E.A. Holm, A.D. Rollett, G.S. Rohrer, Grain boundary energies in

body-centered cubic metals, Acta Materialia 88 (2015) 346-354.

60. K. Linga Murty, I. Charit, Texture development and anisotropic deformation of zircaloys, Prog. Nucl. Energ. 48

(2006) 325-359.

61. F. Emren, U. Von Schlippenbach, K. Lücke, Investigation of the development of the recrystallization textures in

deep drawing steels by ODF analysis, Acta Metall. 34 (1986) 2105-2117.

62. D. Vanderschueren, N. Yoshinaga, K. Koyama, Recrystallisation of Ti IF steel investigated with electron back-

scattering patterns (EBSP), ISIJ Int. 1996 (1996) 1046-1054.

25
List of Tables

Table 1. Analyzed compositions* of four FeCrAl alloys in wt%

Table. 2. EBSD scanning areas for selected specimens

Table 3. Compositions of the matrix and particles in the C36N sample annealed for 8 h at 700 °C, in wt%

26
List of Figures

Figure 1. Recrystallization kinetics of 0.7Nb, 1Nb, and 2Mo in the temperature range of 700-900 °C.

Figure 2. Representative microstructures of as-rolled specimens: (a) IPF map of as-rolled 1Nb; (b) BSE image of

as-rolled 2Nb showing Laves phase particles in bright contrast. Shear bands are marked by arrows in (a). Cracks

within Laves phase particles and the deboning surface between the matrix and the Laves phase are marked by

arrows in (b).

Figure 3. 𝜑2 = 45° ODF map of non-recrystallized areas in 1Nb after annealing for 96 h at 700 °C.

Figure 4. IPF maps of 1Nb annealed at 700 °C for (a) 1 h and (b) 96 h. In (a), boundaries with misorientations

larger than 5° are superimposed as black lines. The average orientations for the A and B areas are (1 0 20)[20 12 -

1] (~ 2.9° from <100>//ND fiber) and (1̅ 1̅ 3)[16 -13 1] (~ 6.5° from α fiber). The point-to-point misorientations

along the dashed lines of L1, L2, and L3 in 1Nb annealed at 700 °C for 96 h are shown in (c) and (d). The IPF

map associated with L3 is superimposed in (d). The origin points are set on the right of L1, L2, and L3.

Figure 5. IPF maps of 1Nb annealed at (a) 800 °C for 24 h, (b) 900 °C for 1 h, and (c) 1100 °C for 0.5 h. Non-

recrystallized areas are indicated by arrows in (a) and (b).

Figure 6. IPF along the ND of recrystallized grains in (a) 2Mo annealed for 1 h at 800 °C, (b) 1Nb annealed for 24

h at 800 °C, and (c) 1Nb annealed for 1 h at 900 °C.

Figure 7. (a) HAADF image of 1Nb after annealing for 8 h at 700 °C. (b-f) Corresponding composition maps of

Fe, Cr, Al, Nb, and Si, respectively.

Figure 8. BSE images of 1Nb after annealing for (a) 96 h at 700 °C and (b) 72 h at 900 °C.

Figure 9. Effect of annealing temperature on the volume fraction of Laves phase in 0.7Nb and 1Nb. For 0.7Nb,

the annealing periods were 96 h (900 °C), 120 h (950 °C), and 48 h (1000 °C), while they were 96 h (700 °C), 72

h (900 °C), 51.5 h (1000 °C), and 24 h (1100 °C) for the 1Nb alloy.

Figure 10. (a) Effect of annealing at 600-900 °C on the Vickers hardness of 1Nb. (b) Comparison of the annealing

effect at 700 °C on the Vickers hardness of 0.7Nb, 1Nb, 2Nb, and 2Mo. (c) Tensile engineering stress vs. strain

27
curves of selected 1Nb specimens. The recrystallization initiation time for studied annealing conditions is

indicated by arrows in (a) and (b).

Figure 11. Taylor factor map with 𝜑2 = 45° (assuming the {110}<111> slip system). Refer to Fig. 3 for

representative orientations in the map.

Figure 12. Area fractions of non-recrystallized <100>//ND fiber, γ fiber, and {112}<110> component in 1Nb after

annealing at various conditions. Annealing conditions are included in the legend with area fractions of

recrystallized grains in parentheses.

Figure 13. Recrystallized grain size of 1Nb (filled markers) and 2Mo (empty markers) after annealing at various

conditions. The error bars represent the standard deviations in the grain size distributions.

28
Table 2. Analyzed compositions* of four FeCrAl alloys in wt%
ID Fe Cr Al Y Mo Si Nb C S O N
0.7Nb 80.19 12.75 6.08 0.040 <0.01 0.19 0.66 0.003 0.0008 0.0009 0.0007
1Nb 79.74 13.11 5.94 0.058 <0.01 0.15 0.98 0.004 0.001 0.0015 0.0003
2Nb 78.82 13.00 5.94 0.055 <0.01 0.20 1.97 0.004 0.001 0.0021 0.0004
2Mo 78.40 13.00 6.29 0.059 1.99 0.20 <0.01 0.001 <0.0003 0.001 0.0004
* Measured by induction coupled plasma optical emission spectroscopy (most elements), combustion analysis (C
and S), and inert gas fusion analysis (O and N).

29
Table. 2. EBSD scanning areas for selected specimens

Alloy ID Annealing conditions Scanning areas (mm2) Scanning locations


1 h/800 °C 2.43 3
0.7Nb 24 h/800 °C 0.81 1
1 h/900 °C 0.81 1
96 h/700 °C 0.86 4
408 h/700 °C 0.89 4
1 h/800 °C 1.04 5
24 h/800 °C 1.02 5
1Nb 72 h/800 °C 0.63 3
214 h/800 °C 0.82 4
1 h/900 °C 0.99 5
24 h/900 °C 0.05 1
30 min/1100 °C 0.71 3
30 min/650 °C 0.19 11
1 h/650 °C 0.16 4
2h/650 °C 0.12 6
2Mo 72 h/650 °C 1.21 1
30 min/750 °C 1.21 1
30 min/800 °C 1.21 1
30 min/900 °C 1.44 1

30
Table 3. Compositions of the matrix and particles in the C36N sample annealed for 8 h at 700 °C, in wt%

Fe Cr Al Y Si Nb
Matrix 81.54 ± 0.73 13.20 ± 0.16 4.87 ± 0.23 0.04 ± 0.05 0.18 ± 0.12 0.17 ± 0.08
Particles 46.72 ± 1.15 5.81 ± 0.25 1.44 ± 0.15 0.02 ± 0.02 3.19 ± 0.14 42.81 ± 1.99

31
Figure 1

32
Figure 2

33
Figure 3

34
Figure 4

35
Figure 5

36
Figure 6

37
Figure 7

38
Figure 8

39
Figure 9

40
Figure 10

41
Figure 11

42
Figure 12

43
Figure 13

44

View publication stats

You might also like