You are on page 1of 9

Journal of Nuclear Materials xxx (2018) xxx-xxx

Contents lists available at ScienceDirect

Journal of Nuclear Materials

F
journal homepage: www.elsevier.com

OO
Effect of thermo-mechanical processing on microstructure and mechanical properties
of U – Nb – Zr alloys: Part 2 - U – 3 wt % Nb – 9 wt % Zr and U – 9 wt% Nb – 3 wt%
Zr
Nathanael Wagner Sales Moraisa, c, Denise Adorno Lopesb, Cláudio Geraldo Schönc, ∗, 1

PR
a
Materials Testing Laboratory, Centro Tecnológico da Marinha, USA
b
College of Engineering and Computing, University of South Carolina, Columbia, SC 29208, USA
c
Department of Metallurgical and Materials Engineering, Escola Politécnica da Universidade de São Paulo, Av. Prof. Mello Moraes, 2463, 05508-030 São Paulo, SP, Brazil

ARTICLE INFO ABSTRACT

ED
Article history: The present work is the second and final part of an extended investigation on U Nb - Zr alloys. It investi-
Received 22 September 2017 gates the effect of mechanical processing routes on microstructure of alloys U - 3 wt % Nb - 9 wt % Zr and
Received in revised form 22 December 2017 U - 9 wt% Nb - 3 wt% Zr, through X-ray diffraction and scanning electron microscopy, completing the inves-
Accepted 24 January 2018
tigation, which started with alloy U - 6 wt% Nb - 6 wt% Zr in part 1. Mechanical properties are determined
Available online xxx
using microhardness and bending tests and correlated with the developed microstructures. The results show
that processing sequence, in particular the inclusion of a 1000 °C heat treatment step, affects significantly the
Keywords: microstructure and mechanical properties of these alloys alloy in different ways. Microstructural characteriza-
CT
Uranium tion shows that both alloys present significant volume fraction of precipitates of a body-centered cubic (BCC)
Uranium alloys γ-Nb-Zr rich phase in addition the uranium-rich matrix. Bending tests show that sample ductility does not cor-
Fuels and fuel elements relate necessarily with hardness and that the key factor appears to be the amount of the γ-Nb-Zr precipitates,
Mechanical properties which controls the matrix microstructure. Samples with a monoclinic α″ cellular microstructure and/or with
Microstructure
the tetragonally-distorted BCC phase (γ0), although not strictly ductile, showed the largest allowed strains-be-
Processing
fore-break and complete elastic recovery of the broken pieces, pointing out to the macroscopic observation of
superelasticity.
RE

© 2017.

1. Introduction chanical properties, in this sense, play a secondary, but important, role
in nuclear fuel design, both during fabrication [4] and during opera-
A recent work by the present authors [1], hereafter called “Part 1”, tion. The evolution of microstructure and mechanical properties dur-
reported the first results of a comprehensive investigation of the effect ing casting, forming and heat treatment operations is, therefore, fun-
R

of processing on microstructure and mechanical properties of 88 wt% damental to predict the behavior of alloy under operational conditions
U – Nb – Zr alloys, focusing there in alloys containing 6 wt% Nb and of irradiation damage.
6 wt% Zr. Two major processing routes, denoted “as cast” and “ho- Radiation is known to decrease fracture toughness of the fuel ma-
CO

mogenized” were investigated as means to produce rolled plates of the terial, both by hardening (i.e by increasing its yield strength due to
alloy, in an attempt to access their viability as metallic nuclear fuel the formation of point defects) and by embrittlement, i.e. increasing its
candidates. ductile-to-brittle transition temperature (DBTT) [5]. The increase of
Uranium alloys are natural candidates to be used as nuclear fuel, the DBTT temperature, approaching the operational temperature, may
due to their naturally high mass density [2]. Current concerns on nu- lead to fracturing in service, enhancing the release of fission gases.
clear safety limit fuel enrichment in new projects to 20 wt % of U235 This increases the inner pressure of the fuel/cladding ensemble [6],
[3] and this higher mass density allows to reach criticality with con- leading to failure of the fuel element.
UN

siderably low enrichment levels. A nuclear fuel element, however, has Strength and ductility of the “as fabricated” fuel are, therefore,
structural functions, in parallel to its neutronic characteristics, and me not only important to understand how the material behaves during
the fuel/cladding assembly, but also to evaluate how they will de-
velop during operation: an initially brittle fuel will likely not be-
∗ come more ductile during irradiation [7], [8]. Understanding the mi-
Corresponding author.
crostructure, therefore, also allows understanding the in service be-
Email addresses: nathanaelmorais@usp.br (N.W. Sales Morais); lopesd@
mailbox.sc.edu (D.A. Lopes); schoen@usp.br (C.G. Schön)
havior of the fuel. These metastable microstructures, naturally, evolve
1
Present address: Department of Materials Science and Engineering, Texas A &
during operation in a reactor core either by action of heat, tend-
M University, 575 Ross St., College Station-TX, 77843 USA. ing towards the thermodynamic

https://doi.org/10.1016/j.jnucmat.2018.01.045
0022-3115/© 2017.
2 Journal of Nuclear Materials xxx (2018) xxx-xxx

equilibrium, or by action of radiation, which disturbs the microstruc- (ICP-OES), is given in Table 1. The nomenclature used in Part 1 will
ture by displacing and/or transmutating atoms. This contributes to fuel be enhanced in the present work to improve results presentation. Dif-
instability and eventual failure [9]. ferent alloys will be referred to using two digits, for instance 39, 66
The use of unalloyed U, although ideal due to the highest possi- and 93. The first digit represents the nominal value of Nb weight frac-
ble density of fissile uranium, is not convenient due to poor mechan- tion and the second the nominal value of Zr weight fraction. The dif-
ical properties and high intrinsic anisotropy. Alloying elements, like ferent sample designations are complemented by up to two letters, the

F
Nb, Zr and Mo, are therefore added to promote the stabilization of first one referring to the processing route (“C” for route A, or “as
the body centered cubic (BCC) phase (γ-U), which has intrinsic good cast”, and “H”, for route B, or homogenized [1]) and the second letter

OO
mechanical and high symmetry in the physical properties [10]. These referring to the particular step within the route (no letter, initial slug,
alloying elements produce additional advantages, for instance Nb is “R”, hot rolled sample, and “A”, hot rolled + annealed sample). Fig. 1
considered also to improve corrosion resistance, strength and ductility reproduce the processing routes, whose details are given in Part 1 [1].
[11], Zr would improve swelling resistance [12] and Mo would im- All samples were quenched in water after each processing step.
prove mechanical properties [13].
Several authors have investigated the effect of additions of Mo, Nb 2.2. Sample characterization
and Zr in U alloys [1,14–20], in particular the thermal stability of the
γ phase have been investigated [21,22]. Alloy processing changes the The procedures for characterization are identical as those described

PR
properties, and this has been mostly investigated in the case of binary in Part 1 [1], they are reproduced here for the sake of completeness.
U – Mo alloys [4,13,23,24]. Fewer works have dedicated to this topic Samples for microstructural characterization were cut with a water re-
in U – Nb – Zr alloys [1,20,25]. It is shown, for example, that the intro- frigerated Al2O3 disc (2500 rpm) in a high precision cutter/grinder. Mi
duction of a homogenization heat treatment intermediate step, is suffi-
cient to eliminate the dendritic microstructure inherited from solidifi-
cation [24]. This dendritic microstructure results in severe banding in Table 1
the hot rolled microstructure and produces more heterogeneous grain Composition of the investigated alloy, determined by ICP-OES.

ED
sizes. In the case of U – Nb – Zr alloys, for example, homogeniza-
Alloy Nb (wt. %) Zr (wt. %) U reference
tion between 700 and 850 °C is used, preceding hot rolling, to obtain a
more homogeneous microstructure [1,20,25]. 39 3.22 ± 0.45 9.22 ± 0.34 balance present work
In Part 1 [1] the present authors investigated the microstructure 66 6.49 ± 0.06 6.63 ± 0.27 balance [1]
and mechanical properties of alloy U – 6 wt% Nb – 6 wt% Zr dur- 93 9.03 ± 0.20 3.22 ± 0.45 balance present work
ing all stages of processing. A complex microstructural evolution was
observed, which was attributed to two competing factors: the damp-
CT
ing of the composition gradients formed by microsegregation in the
“as cast” samples and the changes in volume fraction of Nb – Zr-rich
precipitates. In particular, it was shown, by using bending tests in
miniature-sized sample, that homogenized samples, which presented
the metastable γ phase, or its tetragonal modification, γ0, showed evi-
dences of superelasticity. “As cast” samples, which showed a mixture
RE

of retained γ and the monoclinic martensite α″ showed elastoplastic


behavior in the “as processed” state, suggesting that the presence of
small amounts of the monoclinic martensite is not entirely deleterious
for the alloy's ductility.
The present work concludes the comprehensive examination of the
processing influence on the microstructure and mechanical properties
R

of U – xNb – yZr (with x + y = 12 wt %) using two processing routes,


the first in which the “as-cast” alloy is directly hot rolled to the final
geometry and a second one, in which an intermediate homogenization
step is introduced prior to hot rolling. The present work focuses on al-
CO

loys U – 3 wt % Nb – 9 wt % Zr and U – 9 wt% Nb – 3 wt% Zr. As


previously reported by the present authors [25], the Nb/Zr ratio has
a significant influence on the observed microstructures, following a
trend in which the monoclinic martensite is substituted for the γ0 phase
and then to total retention of γ as this ratio increases. This justifies the
present approach. As in Part 1, the metastable phases will be referred
to using the nomenclature developed by [26], and extended by [27].
UN

2. Experimental procedures

2.1. Materials and sampling

Details on alloy preparation (raw material purity levels, smelting


procedure, alloy surface cleaning, and so on) were identical to the ones
described in Part 1 [1]. Final composition of the alloy slugs, deter-
mined by Inductively Coupled Plasma Optical Emission Spectroscopy
Fig. 1. Synthesis of the processing routes and scheme for sample identification.
Journal of Nuclear Materials xxx (2018) xxx-xxx 3

crostructural analysis was performed via both optical microscopy and


scanning electron microscopy (Phillips® XL – 30) equipped with an
EDAX® EDS detector. Samples were ground and mechanically pol-
ished with 1 μm diamond paste. Final polishing was made by electrop-
olishing with 50% H3PO4 aqueous solution for 60s using 2V potential
and stainless steel as cathode material. The same solution was used

F
for metallographic etching using 5V potential. X-Ray diffraction was
employed using a PANalytical-Empyrean diffractometer with XCel-

OO
erator detector, the measurements were performed at room tempera-
ture with filtered Cu Kα radiation (0.1540598 nm), with 2θ range be-
tween 20° to 80° (the start and end positions respectively) using a
0.003° step. Phase identification and the peak indexation were per-
formed with GSAS [28] software. Ten hardness measurements were
performed in each sample using a Vickers indenter in a semi-auto-
matic Future-Tech durometer with 30 N load and with 15 s holding
time. As explained in Part 1, the determination of the volume fraction

PR
of Nb,Zr-rich particles present in the microstructure is made by image
analysis, using software ImageJ®.
Fig. 2 shows a scheme representing how the different samples
where extracted from the hot rolled plates. It is noteworthy that the po-
sition of the strips used in the bending tests is indicated.

2.3. Bending test

ED
The free bending tests were performed based on standard ASTM
E-290 [29] as in Ref. [1]. The tests were performed in small strips
(16 × 2.5 × 0.7 mm) cut from CR, HR, CA and HA samples in all alloys
(see Fig. 2). The samples were bent in a conventional bench vise. Due
to safety reasons, force was applied with the aid of pliers. Each bend-
ing test was filmed with a steadily attached camera in order to obtain Fig. 3. Definition of the critical fracture angle (εmax) and of the residual angle (εr) in the
CT
the critical fracture angle (εmax) and the residual bending angle of the bending test [1].
two joined broken sample pieces (εr), as defined in Fig. 3. The mea-
surements were accomplished using the last video-frame before the steady increase in volume fraction as processing proceeds, as it was
sample fracture. The measurement of fracture angle was made by an observed for the U - 6 wt% Nb - 6 wt % Zr alloy [1], (these data are
image analysis software. The fracture angles were measured referring also shown in the figure for completeness). The only exception is ob-
from original vertical position of the strip. served for the route A case of the U - 9 wt % Nb - 3 wt % Zr alloy)
RE

which shows a slight decrease of volume fraction as processing pro-


3. Results and discussion ceeds. The volume fractions are also consistently larger for the 66 al-
loy in comparison with either the 39 and the 93 alloys. This is consis-
As shown in Ref. [1], microstructural evolution during processing tent with the postulation of a ternary miscibility gap in the U – Nb –
in alloy U – 6 wt% Nb – 6 wt% Zr can be explained as a consequence Zr system, as suggested in Ref. [1]. The inconsistency between the fi-
of two competing factors: the diffusional damping of composition gra- nal volume fraction in samples 93CA and 93HA can be explained by
R

dients resulting from segregation inherited from solidification, and the microsegregation. Indeed, it is observed that the Nb,Zr-rich particles
growth and coalescence of Nb,Zr-rich particles in the matrix. The pre- tend to disappear in the U-rich regions of sample 93CA, while they
sent analysis, therefore, starts with the evolution of these particles, still grow in the Nb-Zr segregated regions.
which were observed for all three alloys in all processing conditions. Existing information on the phase equilibria in the U – Nb – Zr
CO

Fig. 4 shows the evolution of the Nb,Zr-rich particle volume frac- system is limited to calculated isothermal sections based on ther-
tion as a function of processing for the route A (Fig. 4a) and route modynamic models of the BCC phase. It is well established that
B (Fig. 4b) samples. The results show that these particles present a all three binary systems, U – Nb, U – Zr and Nb – Zr, present
miscibility gaps in the BCC phase [30] [31] [32], [33] calculated
sections at 700 and 750 °C, and, according to the authors, the bi-
nary miscibility gaps extension into the ternary system would be lim-
ited, such that the three miscibility gaps do not interact. The pre-
UN

sent alloys are predicted, according to these calculations, to be sin-


gle-phase γ [34]. on the other hand extrapolates early thermodynamic
models derived for the three binary systems to obtain a model of
the ternary system. In this case, a calculated section at 600 °C pre-
dicts that the BCC miscibility gaps originating from the three bi-
nary systems merge into a three-phase miscibility gap (each phase
rich in one of the components). According to this calculation, the
present alloys should be found in this three-phase field. The exper-
imental results of the present work, as well as those presented in
Fig. 2. Schematic representation of sample extraction from the hot rolled plates. Part 1, are naturally incompatible with both predictions,
4 Journal of Nuclear Materials xxx (2018) xxx-xxx

F
OO
PR
ED
CT
Fig. 4. Evolution of the Nb,Zr-rich particle volume fractions (measured by image analy-
sis) in the samples of route A (a) and route B (b).

showing that a proper experimental examination of the phase equilib-


ria in this system is pertinent.
RE

3.1. Microstructure

3.1.1. Route A
Fig. 5 shows the microstructures of samples 39C (Figs. 5a), 93C
(Fig. 5b) and part of the X-ray diffraction patterns of both alloys (Fig.
R

5c). It is observed that the increased Nb content of alloy 93 leads to the


partial stabilization of the γ-U phase. This is consistent with the ob-
servation of γ-U and the monoclinic martensite (α″) in the 66C sample Fig. 5. Microstructures of samples (a) 39C and (b) 93C, and the corresponding diffrac-
CO

[25]. Still, some residual martensite is found in the 93C sample (the tion patterns (c).
volume fraction is too small to be unambiguously detected in the dif-
fraction pattern, but the accicular morphology of this phase is clearly in the case of the 66C sample, the dendritic morphology of the
observed in Fig. 5b). The Zr-rich alloy (39C) shows also the unequiv- Nb-Zr-rich γ particles, with much finer dendrites, suggest the possibil-
ocal signs of the orthorhombic martensite (α′) in addition to the mon- ity of either a two-liquid equilibrium situation, or of a degenerate eu-
oclinic variant, the two variants seem to be associated with the mi- tectic or peritectic reaction.
crosegregation level in the microstructure, with the orthorhombic vari- Fig. 6 shows the microstructure evolution of alloy U – 3 wt% Nb
UN

ant being observed in the U-richer areas of the microstructure. – 9 wt% Zr in the “as cast” state after rolling (sample 39CR, Fig.
The contrast due to microsegregation in samples 39C and 93C al- 6a) and an additional annealing (sample 39CA, Fig. 6b), as well as
lows a rough estimate the secondary dendrite arm spacing in both al- the corresponding diffraction patterns (Fig. 6c). The results show that
loys, which is evidently smaller in the 39C sample. This estimation is the amount of monoclinic martensite in the microstructure increases
based on the assumption that the contrast originates from different cor- with processing, while the orthorhombic martensite vanishes in sam-
rosion properties of the matrix, which is due to the composition gradi- ple 39CA. The coarsening of the Nb-Zr-rich particles is also evident,
ent existing in the alloy, and, that this gradient is essentially the same consistent with their increased volume fraction.
produced during the solidification of the alloy, such that it reveals a The evolution of the microstructure of alloy U – 9 wt% Nb –
ghost image of the original dendrites. This is consistent with the grain 3 wt% Zr beyond the “as cast” state is shown in Fig. 7. It consists
refining effect of Zr in U – Zr alloys, reported previously by [35]. As of the gradual dissolution of the monoclinic martensite until its to-
tal disappearance
Journal of Nuclear Materials xxx (2018) xxx-xxx 5

F
OO
PR
ED
CT
R RE

Fig. 6. Microstructures of samples (a) 39CR and (b) 39CA, and the corresponding dif- Fig. 7. Microstructures of samples (a) 93CR and (b) 93CA, and the corresponding dif-
CO

fraction patterns (c). fraction patterns (c).

in the 93CA sample (Fig. 7b). The Nb-Zr-rich particles coalesce, as surements in the wider particles). This suggests this phase is α-U, but
in the previous case. Fig. 7b shows fine precipitates in the γ grain a positive identification based on the diffraction pattern is not possible
boundaries. They are too small to positively identify using EDS, but due to the overlap with the peaks attributed to α″ (Fig. 8d).
they appear to be Nb-Zr-rich particles which precipitated in solid state, Fig. 9 shows the diffraction patterns of samples 93H, 93HR and
this could probably account for the apparent reduction of their vol- 93HA and the microstructure of sample 93HA. The images show a
UN

ume fractions, since these small particles are not detected by the image seemingly monophasic microstructure with equiaxed grains. Since the
analysis software. images are virtually identical, only sample 93HA is show, as represen-
tative of the other samples. The diffraction patterns, however, show
3.1.2. Route B that the matrix, evolves from a mixture of γ and γ0 (samples 93H and
The introduction of a homogenization step in alloy U – 3 wt% Nb 93HR) to a pure γ0 state in sample 93HA. As in the case of the U –
– 9 wt% Zr leads to a microstructure dominated by the α″ marten- 6 wt % Nb – 6 wt % Zr alloy [1] this is attributed to the evolution of
site in its cellular morphology (Fig. 8). Samples 39HR and 39HA the γ-Nb, Zr precipitates, which change the chemistry of the matrix. It
(Fig. 8b and c), however, show an additional phase with a bright is important to mention, however, that in previous investigations using
contrast in the Backscattered Electron images, suggesting it is richer a less concentrated alloy [36] [37], the γ0 phase was identified in the
in Uranium, compared with the surrounding matrix (fact corrobo- homogenized samples. This alloy, containing 7.5 wt % Nb and 2.5 wt
rated by EDS mea
6 Journal of Nuclear Materials xxx (2018) xxx-xxx

F
OO
PR
ED
CT
Fig. 9. Microstructure of sample 93HA (a) and the diffraction patterns of samples 93H,
93HR and 93HA (b).

% Zr (Mulberry alloy) shows less propensity for the precipitation of


Nb-Zr-rich precipitates, and, therefore, less freedom to change the
composition of the matrix. In this case the appearance of the γ0 phase
RE

was attributed to residual stresses and to the development of particular


orientations in the recrystallized samples, which would favor the for-
mation of γ0 by a shear mechanism [36].
Table 2 summarizes the lattice parameters obtained for the phases
identified in the diffractograms (only the phases which could be prop-
erly indexed are included). As discussed in Part 1 [1], the lattice para-
R

meters help identifying composition changes in the matrix. For exam


CO

Table 2
Summary of the lattice parameters for the phases identified in the investigated samples
of alloys U – 3 wt % Nb – 9 wt % Zr and U – 9 wt % Nb – 3 wt % Zr.

Sample Phase a0 [nm] b0 [nm] c0 [nm] Other

39C α′ 0.2887 0.5816 0.4992


39CR α″ 0.2927 0.5914 0.5077 γ = 91.3°
39CA α″ 0.2940 0.5885 0.5020 γ = 92.2°
UN

39H α″ 0.2874 0.5773 0.5005 γ = 91.5°


39HR α″ 0.2825 0.5865 0.5056 γ = 93.1°
39HA α″ 0.2853 0.5924 0.5056 γ = 93.2°
93C γ 0.3468
α″ 0.2864 0.5740 0.4895 γ = 91.4°
93CR γ 0.3558
α″ 0.2877 0.5693 0.5143 γ = 92.3°
93CA γ 0.3580
Fig. 8. Microstructures of samples (a) 39H, (b) 39HR and (c) 39HA, and the corre- 93H γ 0.3472
sponding diffraction patterns (d). 93HR γ 0.3483
γ0 0.5052 0.3502 c/a = 0.693
93HA γ 0.3543
γ0 0.5053 0.3579 c/a = 0.708
Journal of Nuclear Materials xxx (2018) xxx-xxx 7

ple, increased Nb levels in the α″ martensite are associated with an in-


crease in the γ angle [38], while an increase in a0 for the γ phase could
indicate increased levels of dissolved Nb and Zr.
Direct comparison with the literature is possible only in similar
systems. [35], for example, investigated the α′ martensite in U – Zr
alloys, providing a description of the lattice parameter evolution as a

F
function of composition for this phase. Applying this functional de-
pendency to a hypothetical U – 12 wt% Zr alloy (= 26.1 at.% Zr) re-

OO
sults in a0 = 0.2854 nm, b0 = 0.5860 nm and c0 = 0.4935 nm, which com-
pared with the data for sample 39C shows that substitution of Zr
by Nb leads to an expansion along directions [100] and [001] and
a contraction along [010], in general the predicted atomic volume
of 2.064 × 10−2 nm3 is increased to 2.095 ×10 − 2nm3 with this substi-
tution. This phase was detected in the diffraction pattern of sample
39CR, but the obtained lattice parameters are inconsistent with the re-
ported values, probably due to the weak signal and to the small vol-

PR
ume fraction, therefore they were not included in Table 2.
Phases α″ and γ0 were investigated by [39] in Mulberry's alloy
and by [38] in U – Nb alloys. The first author reports a0 = 0.289 nm,
b0 = 0.568, c0 = 0.499 and γ = 91° for α″ and a0 = 0.4948 nm,
c0 = 0.3371 nm (c/a = 0.681) for γ0. Both values are consistent with the
ones reported in Table 2 [38]. reports parameters a0 = 0.285 nm, b0
from 0.581 to 0.577 nm, c0 = 0.499 to 0.500 nm and γ from 90.8 to 91.5
for alpha″ and a0 from 0.697 to 0.694 nm and c0 = 0.338 nm for γ0 (c/a≈

ED
0.49). This shows that substitution of Nb by Zr leads to an expansion
of a0 and c0 and a contraction of b0 (as well as in an increase in angle
γ), which had already been reported by [39] in the Mulberry alloys.

3.2. Mechanical properties


CT
Fig. 10 shows the hardness evolution of the samples of the three Fig. 10. Hardness evolution in alloys 39, 66, and 93, as a function of processing, for
alloys in the different phases of processing, for the two processing processing route A (a) and B (b).
routes. The results of alloy U – 6 wt % Nb – 6 wt % Zr are taken
from Ref. [1] and are reproduced for the sake of completeness. As ex- “austenite” phase in the present case, so the observation of superelas-
pected, sample which present either the α′ or the α″ martensitic micro- ticity would be linked to the stress-induced formation of α″. Super-
RE

constituents are considerably harder in comparison with the samples elasticity, however, can be observed in martensitic structures in some
in which only γ or γ0 phases are present. This tetragonal modification shape memory alloys [40], linked the reversible movement of twinn
of the γ phase seems to be slightly harder then γ, though. These figures boundary variants in the martensite. This could justify the full re-
also allow to draw some conclusions about the roles of Nb and Zr in versibility of the samples which present a fully transformed α″ mi-
phase stability of these Uranium alloys. In fact, it is observed that a crostructure.
tendency for stabilization of γ (or γ0) exist with increasing Nb content.
R

It is further concluded that the Nb content of alloy U – 3 wt % Nb – 3.3. Synthesis


9 wt % Zr is insufficient to prevent the precipitation of α-U during the
processing stages in the homogenized samples. The fuel core must be as plastic as possible in order to accommo-
As discussed previously, ductility in Uranium alloys cannot be dis- date the stresses from forming, the stresses during reactor assembly
CO

cussed only based on hardness. Fig. 11 shows the maximum and resid- and those observed in initial service. If we consider only this criterion,
ual bending angles measured for both alloys as a function of pro- the best results, both in maximum plasticity (φr) and in minimum hard-
cessing. Sample 39CR shows larger bending angles before fracture in ness, among all investigated alloys are obtained through route A. Con-
comparison with sample 39HR, while the bending angles of samples sidering the results obtained in Part 1 [1] as well as those presented
39CA and 39HA is quite similar, a behavior which was also observed in the present work, the processing recommendation to eventually ob-
in samples 66HR and 66HA [1]. As discussed in Part 1, this could be a tain a fuel alloy, therefore, is to perform the hot rolling without previ-
consequence of the role of the Nb-Zr-rich particles in fracture. It must ous homogenization heat treatment followed by the annealing/quench-
UN

be remembered, however, that the volume fraction of these particles is ing heat treatment (in particular in alloys 66CA and 93CA). It must
smaller in both alloys for this processing route, in comparison with the be noticed, however, that the route B samples, in particular 93HR and
U – 6 wt % Nb - 6 wt % Zr case, and the variation of volume fraction 93HA show superior maximum bending angles, even considering they
with processing is also less severe. are superelastic (that is, no residual plasticity is observed). This does
As in the case of the U - 6 wt %Nb – 6 wt % Zr alloy, some samples not constitute, therefore, proper ductility, but these alloys are certainly
showed complete elastic recovery after fracture, showing macroscopic prone to survive severe loads during fuel assembly, without fracturing.
evidence of superelasticity. It is interesting to observe that this phe- This, of course, does not guarantee that the fuel core will re-
nomenon is observed in samples containing either α″ or γ + γ0 in the main plastic in long-term service. Exposition to high temperatures
matrix. It is supposed that the γ, or perhaps the γ0 phase is the and irradiation will cause microstructural evolution and property de-
grading [5]. Based on the irradiation experiment conducted by [41],
however, high
8 Journal of Nuclear Materials xxx (2018) xxx-xxx

BCC precipitates, which alter the composition of the matrix as pro-


cessing proceeds.
With regard to the mechanical properties, hardness correlates with
the microstructure, increasing when the matrix contains α″ and de-
creasing as γ dominates the matrix. The tetragonally-distorted BCC
modification (γ0) slightly increases hardness. Using the bending tests,

F
however, it was shown that superelasticity identified in the U – 6 wt
% Nb – 6 wt % Zr alloy in γ + γ0 matrices (Part 1), is possible also for

OO
alloys which present α″, suggesting a thermo-elastic behavior of this
martensite.
Concerning the potential use as fuel material, considering both the
results of the present work and of Part 1, it is concluded that the opti-
mal combination of plasticity, either measured by bending or by hard-
ness, and stabilization of the γ phase is obtained for alloy U – 9 wt
% Nb – 3 wt % Zr, rolled directly from the “as cast” state. The intro-
duction of an annealing step after rolling, although not compromising,

PR
seems to be unnecessary. In particular, the realization of a homoge-
nization step prior to rolling is avoidable, since it has little effect over
the properties.

Acknowledgments

This work has been financially supported by the Brazilian National

ED
Research Council (CNPq, Brasília-DF, Brazil) under Proj. 312424/
2013-2. The support by the Brazilian Navy, in the form of materials
and grants, is gratefully appreciated.

Appendix A. Supplementary data


CT
Supplementary data related to this article can be found at https://
doi.org/10.1016/j.jnucmat.2018.01.045.
Fig. 11. Bending angles for alloys 39 (a) and 93 (b).

plasticity is important in order to survive the swelling caused by the References


fission gases bubbles. Although it is not possible to conclude that an
initially plastic alloy will remain plastic under irradiation, it is reason- [1] N.W.S. Morais, D.A. Lopes, C.G. Schön, J. Nucl. Mater. 488 (2017) 173–180.
RE

able to consider that the highest initial plasticity will lead to better in [2] A.N. Holden (Ed.), Physical Metallurgy of Uranium, Addison-Wesley Publish-
service plasticity. Based on mechanical and microstructural data ob- ing Co., Geneve, Switzerland, 1958.
[3] S.J. Chilk, Federal Registry, 51, 1986, URL http://www.rertr.anl.gov/
tained in the present works, and the observation that the increase in REFDOCS/NRCRULE.pdf, Accessed 15 March 2016.
niobium content seems to improve the stability of the γ phase, alloy U [4] D.E. Burkes, R. Prabhakaran, T. Hartmann, J.-F. Jue, F.J. Rice, Nuclear Eng.
– 9 wt % Nb – 3 wt % Zr either in the CR or in the CA condition (that Design 240 (2010) 1332–1339.
is either 93CR or 93CA) is a good candidate to be used as fuel for re- [5] G.S. Was, Fundamentals of Radiation Materials Science, Springer Verlag, Berlin
Heidelberg, 2010.
R

search reactors. [6] D.E. Burkes, D.J. Senor, A.M. Casella, Nuclear Eng. Design 310 (2016) 48–56.
[7] J.E. Gates, E.G. Bodine, J.C. Bell, A.A. Bauer, G.D. Calkins, Stress-strain Prop-
4. Conclusions erties of Irradated Uranium - 10 W/o Molybdenum, Technical Report
BMI-APDA-638, Batelle Memorial Institute, Columbus-OH, 1958https://doi.
CO

org/10.2172/4312625.
The present work completes the systematic study of U – Nb – Zr al- [8] B.W. Chung, K.E. Lema, P.G. Allen, J. Nuclear Mater 471 (2016) 239–242.
loys containing 88 wt % U, initiated in Part 1, concerning the mechan- [9] T.A. Pedrosa, A.M.M. Santos, F.S. Lameiras, P.R. Cetlin, W.B. Ferraz, J. Nu-
ical properties and microstructures developed as a function of process- clear Mater 457 (2015) 100–117.
ing. The results show that processing, in particular the introduction of [10] A.E. Dwight, M.H. Mueller, Constitution of the Uranium-rich U – Nb and U –
a homogenizations step prior to hot rolling, has a strong impact in the Nb – Zr, Technical Report Argonne National Laboratory, 1957, ANL-5581.
[11] H.M. Volz, R.E. Hackenberg, A.M. Kelly, W.L. Hults, A.C. Lawson, R.D. Field,
obtained microstructures. D.F. Teter, D.J. Thoma, J. Alloys Compounds 444-445 (2007) 217–225.
Alloy U – 3 wt % Nb – 9 wt % Zr shows monoclinic martensite (
UN

[12] G. Lagerberg, J. Nucl. Mater. 9 (1963) 261–276.


α″) in the matrix, both in the “as cast” state and in the homogenized [13] D.E. Burkes, R. Prabhakaran, J.-F. Jue, F.J. Rice, Metall. Mater. Trans.
state. The homogenized samples present, both after hot rolling and af- 40 (2009) 1069–1079.
[14] J.-M. Park, K.-H. Kim, D.-S. Sohn, C.-K. Kim, G.L. Hofman, J. Nucl. Mater.
ter hot rolling followed by annealing, an uranium-rich phase which is 265 (1999) 38–43, https://doi.org/10.1016/S0022-3115(98)00647-3 http://www.
believed to be α-U. sciencedirect.com/science/article/pii/S0022311598006473.
Alloy U – 9 wt % Nb – 3 wt % Zr, on the other hand, develops a [15] M.K. Meyer, G.L. Hofman, T.C. Wiencek, S.L. Hayes, J.L. Snelgrove, J. Nucl.
stable γ matrix in the “as cast” state, which is replaced by the tetrago- Mater. 299 (2001) 175–179.
nally-distorted γ0 variant in the homogenized state. Key factor in these
developments appears to be the growth and coalescence of Nb-Zr-rich
Journal of Nuclear Materials xxx (2018) xxx-xxx 9

[16] C.-T. Lee, J.-H. Park, T.-K. Kim, U.-J. Lee, B.-S. Lee, D.-S. Sohn, J. Nucl. [28] A.C. Larson, R.B. Von Dreele, General Structure Analysis System (GSAS), Los
Mater. 373 (2008) 275–279, https://doi.org/10.1016/j.jnucmat.2007.06.006 http: Alamos National Laboratory, 2004, Technical Report LAUR 86–748.
//www.sciencedirect.com/science/article/pii/S002231150700832X. [29] A. S, E290, Standard Test Methods for Bend Testing of Materials for Ductility,
[17] D.A. Lopes, T.A.G. Restivo, A.F. Padilha, J. Nucl. Mater. 440 (2013) 304–309. 2014https://doi.org/10.1520/E0290-14.
[18] S. Van der Berghe, P. Lemoine, Nucl. Eng. Tech 46 (2014) 125–146. [30] A. Fernandez-Guillermet, Z. Matallk. 82 (1991) 478–487.
[19] K. Ghoshal, T. Kutty, S. Mishra, A. Kumar, J. Nucl. Mater. 432 (2013) 20–22, [31] S. Ahn, S. Irukuvarghula, S.M. McDeavitt, J. Alloys Comp 611 (2014) 355–362.
https://doi.org/10.1016/j.jnucmat.2012.08.007 http://www.sciencedirect.com/ [32] T. Duong, R.E. Hackenberg, A. Landa, P. Honarmandi, A. Talapatra, H.M.

F
science/article/pii/S0022311512004163. Voltz, A. Llobet, A.I. Smith, G. King, S. Bajaj, A. Ruban, L. Vitos, P.E.A.
[20] K. Ghoshal, S. Kaity, S. Mishra, A. Kumar, J. Nucl. Mater. 446 (2014) 217–226. Turchi, R. Arróyave, Calphad 55 (2016) 219–230.
[21] L.I. Gomozov, V.A. Makhova, V.V. Gomov, O.S. Ivanov, Fiz. Metal. Metallov [33] A.L. Udovskij, O.S. Ivanov, In: Thermodynamics of Nuclear Materials, vol. 2,

OO
33 (1972) 307–314. International Atomic Energy Agency, Vienna, 1974, pp. 285–302.
[22] C.L. Komar Varela, L.M. Gribaudo, R.O. González, S.F. Aricó, J. Nucl. Mater. [34] R.O. Williams, J. Nucl. Mater. 82 (1979) 184–192.
453 (2014) 124–130. [35] C. Basak, G.J. Prasad, H.S. Kamath, N. Prabhu, J. Alloys Compounds
[23] V.V. Joshi, E.A. Nyberg, C.A. Lavender, D. Paxton, H. Garmestani, D.E. 480 (2009) 857–862.
Burkes, J. Nucl. Mater. 465 (2015a) 805–813. [36] D.A. Lopes, T.A.G. Restivo, N.B. Lima, A.F. Padilha, J. Nucl. Mater.
[24] V.V. Joshi, E.A. Nyberg, C.A. Lavender, D. Paxton, D.E. Burkes, J. Nucl. 449 (2014) 23–30.
Mater. 465 (2015b) 710–718. [37] D.A. Lopes, A.J.O. Zimmermann, T.A.G. Restivo, A.F. Padilha, Metal. Mater.
[25] N.W.S. Morais, M.A. Tunes, V.O. Santos, C.G. Schön, R.G. Gomide, In: Top Trans E 2 (2015) 147–156.
Fuel 2015–Conference Procedings Poster, European Nuclear Society, Brussels, [38] K. Tangri, D.K. Chaudhuri, J. Nuclear Mater 15 (1965) 278–284.

PR
Belgium, 2015, pp. 403–413 http://www.euronuclear.org/events/topfuel/ [39] H.L. Yakel, J. Nuclear Mater 33 (1969) 286–295.
topfuel2015/transactions/topfuel2015-transactions-poster.pdf, Accessed 18 [40] T. Tadaki, K. Otsuka, K. Shimizu, Ann. Rev. Mater. Sci. 18 (1988) 25–45.
March 2016. [41] J. Gan, B.D. Miller, D.D. Keiser Jr., J.F. Jue, J. Madden, A.B. Robinson, H.
[26] J. Lehmann, R.F. Hills, J. Nuclear Mater 2 (1960) 261–268. Ozaltun, G. Moore, M.K. Meyer, J. Nucl. Mater. 492 (2017) 195–203, https://
[27] K. Tangri, Mem. Sci. Rev. Metall. 58 (1961) 469–477. doi.org/10.1016/j.jnucmat.2017.05.035.

ED
CT
R RE
CO
UN

You might also like