Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Nuclear Materials 447 (2014) 150–159

Contents lists available at ScienceDirect

Journal of Nuclear Materials


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

Fabrication of simulated plate fuel elements: Defining role of stress relief


annealing
D. Kohli a,b,⇑, R. Rakesh a, V.P. Sinha a, G.J. Prasad a, I. Samajdar b
a
Metallic Fuels Division, Bhabha Atomic Research Centre, Trombay, Mumbai 400085, India
b
Department of Metallurgical Engineering and Materials Science, IIT Bombay, Powai, Mumbai 400076, India

a r t i c l e i n f o a b s t r a c t

Article history: This study involved fabrication of simulated plate fuel elements. Uranium silicide of actual fuel elements
Received 27 July 2013 was replaced with yttria. The fabrication stages were otherwise identical. The final cold rolled and/or
Accepted 1 January 2014 straightened plates, without stress relief, showed an inverse relationship between bond strength and
Available online 10 January 2014
out of plane residual shear stress (s13). Stress relief of s13 was conducted over a range of tempera-
tures/times (200–500 °C and 15–240 min) and led to corresponding improvements in bond strength.
Fastest s13 relief was obtained through 300 °C annealing. Elimination of microscopic shear bands, through
recovery and partial recrystallization, was clearly the most effective mechanism of relieving s13.
Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction An earlier study [13], as well as the present one, explored


fabrication of simulated plate fuel elements. Yttria (Y2O3) replaced
Plate fuel elements are widely used in high flux research uranium silicide as surrogate fuel in a typical (Fig. 1a) picture-
reactors involved in the production of radioisotopes, irradiation frame assembly. The fabrication stages (Fig. 1b) were otherwise
studies, etc. [1–7]. Roll bonding is the usual fabrication route identical. The preceding study [13], by the same team, had docu-
[1,3–8]. The final cold rolled plate is expected to have high stress mented aspects of microstructural developments at all (except
gradients. This has been shown through simulations [9–11], as well stress relief) fabrication stages. This study, on the other hand, tries
as with experimental observations on plate fuels [12,13] and also to rationalize careful observations on residual stresses with devel-
in accumulative roll bonding [14,15]. A preceding study [13] has opments in microstructures and bond strength.
shown that such stress gradients are mainly in terms of out of
plane residual shear stress (s13). A direct correlation between s13
and bond strength between, aluminum cladding and simulated 2. Experimental details
fuel meat (of aluminum and yttria particles), was also indicated.
The stress relief process, typically given at the end of fabrication, Fig. 1a shows the schematic of a picture-frame assembly. Com-
is a topic of considerable academic/applied interests [9,16–18]. Be- position of AA6061 cladding plates is given in Table 1. Surrogate
cause of its apparent correlation with bond strength [12,13], stress fuel meat consisted of a compact: aluminum and yttria (Y2O3) par-
relief is naturally of concern to the fabrication of plate fuel ele- ticles (60:40 by volume) of 20–100 lm size. This was cold com-
ments. However, comprehensive studies relating stress relief with pacted at 900 MPa stress and a density of P95%. Y2O3 is selected
property (such as bond strength) changes and microstructural as surrogate fuel instead of actual uranium silicide (U3Si2). The
developments are virtually non-existent. It was hence decided to two materials have similar elastic modulus and coefficient of ther-
initiate the present study: role of stress relief on bond strength mal expansions. It may also be noted that a preceding study [13], in
and microstructural developments. the same journal, used this surrogate fuel and made observations
on plastic deformation. The present manuscript, on the other hand,
focuses on the stress relief annealing: the final stage of plate fuel
Abbreviations: AA, Aluminum Association; ASTM, American Society for Testing fabrication. Fig. 1b shows the process flow sheet used. Readers
and Materials; EBSD, Electron Backscattered Diffraction; RD, Rolling Direction; ND, may find further description in Ref. [13]. The final cold rolling
Normal Direction; TD, Transverse Direction; GIXRD, Grazing Incidence X-Ray was often followed by industrial straightening operation [18].
Diffraction; XEC, X-ray Elastic Constant; FEG, Field Emission Gun; SEM, Scanning
The straightening consisted of 10 set of rolls placed on a rolling ta-
Electron Microscope; OIM, Orientation Imaging Microscopy.
⇑ Corresponding author at: Metallic Fuels Division, Bhabha Atomic Research ble. This provided cyclic, albeit alternating, bending but did not im-
Centre, Trombay, Mumbai 400085, India. Tel.: +91 22 25593929. pose reductions in thickness. The final plates were subjected to
E-mail address: divyak@barc.gov.in (D. Kohli). stress relief (followed by furnace cooling with similar cooling rates

0022-3115/$ - see front matter Ó 2014 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jnucmat.2014.01.002
D. Kohli et al. / Journal of Nuclear Materials 447 (2014) 150–159 151

(a)

(b)
Fig. 1. (a) Schematic showing typical picture-frame assembly. This included aluminum (AA6061) cladding (cover plate and picture frame) and surrogate fuel meat. The latter
consisted of 60:40 (by volume) aluminum and yttria (Y2O3) cold compacted powder. (b) Process flow sheet for simulated plate fuel fabrication. In the actual plate fuel
fabrication, Y2O3 is replaced with uranium silicide (U3Si2). Process flow sheet is otherwise identical.

Table 1
Composition of AA6061 in wt.% alloying elements. quarter in all the dimensions. At least three samples were mea-
sured. Limited numbers of samples were also subjected to pull test
Mg Si Mn Fe Cu Cr Zn Ti Al
[20]. This established, also shown in Ref. [13], a similar trend be-
1 0.67 0.19 0.20 0.30 0.024 0.015 0.022 97.57 tween peel and pull test data.
Characterization involved residual stress and EBSD measure-
ments. Stress measurements were made with appropriate depths
of 0.7 °C/min) at various temperatures (200 °C, 300 °C, 400 °C and of X-ray penetration (to be discussed latter) of the rolling plane
500 °C) and times (0, 15, 30, 60 and 240 min). (plane containing RD and TD). On the other hand, EBSD scans were
Mid-width sections of the rolled plates, with and without stress taken on the long-transverse section (containing RD and ND). EBSD
relief, were taken for peel tests. This was carried out with ASTM samples were prepared by standard metallography followed by
D1876-01 standard [19], after proportionately scaling down one electropolishing. Electropolishing involved an electrolyte of
152 D. Kohli et al. / Journal of Nuclear Materials 447 (2014) 150–159

perchloric acid and methanol (20:80 by volume), 20 °C and at x = incident/grazing angle between the incident beam and
13 V dc. Controlled polishing with sub-micron colloidal silica was specimen surface.
used for residual stress sample preparation. It was confirmed that
electropolishing and colloidal silica polish returned similar values In GIXRD, reflections for different {h k l} or Bragg angles (h) are
of residual stresses. This eliminated possibilities of stress relaxa- obtained. This provides different w:
tion through controlled colloidal silica polishing.
Residual stress measurements were made on a Panalytical™ Wh k l ¼ hh k l  x ð1Þ
X’Pert PRO MRD system. A commercial software, X’Pert Stress
Residual stress measurements estimate appropriate interplanar
Plus™, was used. Fig. 2a and b shows the diffraction geometry hkl
spacings (d/w ):
for GIXRD. It is important to define:
/ = rotation around the specimen plane normal. hkl o
d/;w  d
w = inclination between specimen plane normal and the diffrac- eh/;wk l ¼ o
tion vector. d

Fig. 2. (a) Diffraction geometry for grazing incidence X-ray diffraction (GIXRD). (b) Definitions of w and /: details deliberated in the text. Also included are: angular
conventions of rotations between laboratory (L) and specimen (S) co-ordinate systems and conventions for tensorial components of the residual stress. (c) The GIXRD profiles
at different grazing angles (x). (d) Estimated s13 at different distances from the interface. Standard deviations of stress measurements are shown as appropriate error bars.
D. Kohli et al. / Journal of Nuclear Materials 447 (2014) 150–159 153

Fig. 3. (a) Images from simulated plate fuel fabrication: with plates undergoing successful bonding or severe debonding. Multiple {h k l} GIXRD estimated (b) r11 and (c) s13 at
cladding-surrogate fuel interface versus bond strength. Data include plates from final fabrication (without stress relief): with and without post-fabrication straightening
operation. Standard deviations of stress measurements are shown as appropriate error bars.

Fig. 4. (a) Bond strength with annealing time and temperature. Error bars indicate standard deviations from multiple peel-test measurements. Also included are
representative fractographs from samples with low and high bond strengths. (b) Rate of change of bond strength (0–30 min annealing) with annealing temperature.
154 D. Kohli et al. / Journal of Nuclear Materials 447 (2014) 150–159


1  2

2 By using a long incident beam path for small x, the above equation
¼ S1 ðr11 þ r22 þ r33 Þ þ S2 r11 cos2 / þ r22 sin / þ s12 sin2/ sin w
2 can be simplified as

1 1 sin x
þ S2 r33 cos2 w þ S2 ðs13 cos/ þ s23 sin/Þ sin2w ð2Þ C¼ ð5aÞ
2 2 l
rij and sij represent normal and shear components of a stress matrix In this l is termed as linear absorption coefficient and was taken as
(Fig. 2b). S1 and S2 are XEC, which can be related to the Poisson’s ra- 135.621 cm1 [29]. Controlled colloidal silica polish plus variable x,
tion (t) and Young’s modulus (E): allowed a calibration for C. This is shown in Fig. 2c. At x = 3.5°,
t Y2O3 peak was observed. Microscopic and estimated (Eq. (5)) values
S1 ¼  ð3Þ
E of C were approximately 4.5 lm cladding thickness. Such samples,
with 4.5 ± 0.5 lm cladding thickness from the cladding-surrogate
1þt meat interface, were used for subsequent stress relief annealing.
S2 ¼ ð4Þ
E For dislocation density estimation XRD system with line focus
E = 69.8 GPa was measured experimentally for the AA6061. For and high resolution (0.01° step size) h2h (where h is the Bragg an-
t = 0.34, the value from literature [21] was taken. gle) scans were used. The analysis was carried out in accordance
Standard d-sin2w measurements [22–25] involve a range of w with the 4th order variance method [30–33]. The estimated dislo-
values. In GIXRD [22,26–28], on the other hand, different {h k l} of- cation densities, at each point, were normalized with prior defor-
fer a range of d and w (Eq. (1)). A linear d-sin2w was then plotted mation values. EBSD measurements were done in FEI™ Quanta
with appropriate extrapolation [23,24] for w-splitting. 3D-FEG SEM using TSL-OIM™ system. Step size of 0.1 lm was used
This study used uniaxial stress analysis. Sample RD was aligned with identical beam and video conditions for all the scans. The
for u = 0°: allowing measurements of r11 and s13. 1, 2 and 3 EBSD scans were taken before and after stress relief annealing,
respectively represent the RD, TD and ND of the specimen using a suitably methodology described elsewhere [34].
(Fig. 2b). For peak identification Pearson VII profile fitting was
used, with appropriate user inputs for ‘minor’ shifts. In GIXRD, con- 3. Results
trolling x can offer different depths of penetration (C according to
Eq. (5) [26]: Fig. 2d plots s13 (out of plane residual shear stress) as a function
 1 of distance from cladding-surrogate meat interface. It is apparent
l l that a significant s13 gradient existed in the cold rolled plate. Near
C¼ þ ð5Þ
sin x sinð2hh k l  xÞ interface s13 was the highest. This study tried, with controlled

Fig. 5. (a) s13 and bond strength as function of annealing time for (a) 200°, (b) 300°, (c) 400° and (d) 500 °C annealing. Appropriate error bars, indicating standard deviations,
are included. s13, residual shear stress at cladding-meat interface, were obtained by GIXRD at 4.5 ± 0.5 lm cladding thickness from the cladding-surrogate meat interface.
D. Kohli et al. / Journal of Nuclear Materials 447 (2014) 150–159 155

colloidal silica etch-polish (described earlier), to estimate near constant, Q is the activation energy in eV, C is material constant and
t;T
interface s13 for a cladding depth of 4.5 ± 0.5 lm. Straightening at m is ‘stress relaxation exponent’. A plot of log (ln (st¼0
13 =s13 ) versus
times led to complete debonding and almost always to a drop in log time (Fig. 6b) gives the value of stress relaxation exponent
bond strength. Extreme example of the former is shown in (m). Fig. 6c shows that m and rate of increase in bond strength
Fig. 3a. Though estimated average bond strength did not relate to had a similar scaling with annealing temperatures. Both had a ‘high-
r11 (Fig. 3b), Fig. 3c demonstrates an inverse relationship between est’ at 300 °C.
bond strength and s13. This validates a critical observation of the To understand the rationale behind best stress relief kinetics at
earlier study [13]: the loss of bond strength through developments 300 °C, direct EBSD measurements were made, see Fig. 7. These re-
of near interface out of plane residual shear stress (s13). vealed temperature dependent (for 30 min annealing) develop-
Fig. 4a collates the evolution of bond strength with stress relief. ments in annealed microstructures. Annealing at 200 °C did not
Stress relief clearly improved the bond strength: nature of fracture significantly affect the deformed grain structure, while partial
surface changed from brittle-ductile to fully ductile. Though bond recrystallization was noted after 300 °C stress relief. On the other
strength values were not remarkably different beyond 60 min hand, 400 and 500 °C annealing led to an ‘apparent’ full recrystal-
stress relief, initial 30 min clearly had different rates of bond lization. These observations raised an interesting point. They had
strength improvements – see Fig. 4b. The fastest improvement shown that partial recrystallization was more effective than an
was at 300° stress relief. This appears counter-intuitive. ‘apparent’ full recrystallization. A rationale for this counter-intui-
Fig. 5 shows effects of stress relief at different annealing tem- tive observation is sought in the subsequent section on discussion.
peratures. It is clear that drops in s13 matched corresponding
improvements in the bond strength. Fig. 6a confirms, for the initial
30 min annealing, that the rate of improvement in bond strength
has a direct correlation with rate of relaxation of residual shear 4. Discussion
stress (s13). This was further analyzed with Zener–Wert–Avrami
function [35–37]: The present study brought clear observations on temperature
dependent stress relief. Before initiating discussions on microstruc-
   m 
st;T
13 Q tural developments and stress relief, it is important to reiterate
¼ exp  C  t  exp  ð6Þ observations on improved bond strength. Rate of change in bond
s13
t¼0 kT
strength, for initial 30 min annealing, was best at 300 °C (Fig. 4b).
t;T
where s13 is the residual stresses at time t (in s) and temperature T This was clear and statistically reproducible. The only plausible
(in K), s13 is the initial residual stresses in MPa, k is the Boltzmann
t¼0
explanation is non-monotonic (with annealing temperature)
156 D. Kohli et al. / Journal of Nuclear Materials 447 (2014) 150–159

Before Stress Relief After Stress Relief


(30 min.)

200ºC

RD

TD 300ºC

400ºC

500ºC

Fig. 7. Direct observations on the microstructure developments during stress relief (30 min annealing). IPF (inverse pole figure) images (with data points below 0.1 CI marked
black) of identical regions before and after the stress relief treatments. 300 °C stress relief showed partial recrystallization, while microstructures above 300 °C appeared fully
recrystallized.

changes in residual stresses and/or effective diffusivities. The effec- (Fig. 4b) after 300 °C annealing. Hence the experimental observa-
tive diffusivity [38,39] can be expressed as: tion on best bond strength after 300 °C annealing can only be
explained from non-monotonic temperature dependence of
d
Deff ¼ Dgb þ gDpipe þ Dl ð7Þ residual stresses and stress relief.
d
This study established, see Fig. 3, that s13 developments were
  primarily through cold working. Kinetics of subsequent stress relief
Q
D ¼ D0 exp  ð8Þ (Fig. 6a) was clearly temperature dependent. Intermediate anneal-
RT
ing temperature of 300 °C offered fastest s13 relief and bond
where Dgb = grain boundary diffusivity, Dpipe = pipe (dislocation) dif- strength improvement. Arguably, high temperature annealing
fusivity, Dl = lattice diffusivity, d = grain size, d = grain boundary may offer better diffusivity and corresponding improvements in
width and g = cross-sectional area of dislocations per unit area of metallurgical bonding. However, this also have the potential of cre-
matrix. Eq. (8) represents temperature (T) dependence of diffusivi- ating residual stresses through thermal mismatch. This study used
ties, Q being the activation energy for diffusion. very slow furnace cooling (approximately 0.7 °C per minute),
Annealing temperature/time brought in changes in grain size which is expected to reduce (but not eliminate) thermal stresses.
d (Fig. 8a) and in dislocation densities (Fig. 8b). These data were The temperature dependence of the stress relief annealing
combined with other constants (see Table 2) and suitable mea- (Fig. 6a) was, however, apparent irrespective of the cooling
sures of d and g. The value of d was taken [38] as 0.5 nm while profiles.
g was calculated considering 10 atoms in a dislocation cross- An effort has been made to explain these observations
section with matrix containing 1013 atoms/mm2 [38]. Estimated through a hypothesis and complimentary phenomenological evi-
effective diffusivities, as summarized in Fig. 8c, shows clear dences, see Fig. 9. As shown in Fig. 9a(i), during unconstraint
temperature dependence. Annealing at 500 or 400 °C had 3 deformation of cladding and surrogate fuel meat: higher strain
and 2 orders of magnitude higher diffusivities than 300 °C. in cladding is expected. It is recorded in [13] that initial hot
Possible differences in diffusivities and diffusion bonding thus working offered significant difference in the strain partitioning.
cannot explain highest rate of increase in bond strength Strain in the cladding was more. Subsequent deformations
D. Kohli et al. / Journal of Nuclear Materials 447 (2014) 150–159 157

Fig. 8. (a) Grain size, (b) normalized dislocation density (estimated dislocation density normalized by the value in fully recrystallized state), (c) estimated effective diffusivity
(Eq. (7)) are plotted as a function of annealing time and temperature. (a) Contains standard deviations in grain size from multiple measurements.

Table 2
is shown through direct EBSD observations (Fig. 9b). This was
Values of constants (Eqs. (7) and (8)) used for estimating effective diffusivities. clearly the most effective s13 relief.
Initial stages (30 min) of 400 and 500 °C annealing led to the
D0 (cm2/sec) Q (kcal/mol)
formation of large grains, see Fig. 7. Visually they appeared fully
Grain boundary diffusion [40] 0.1 14.4 recrystallized. Such grains can originate through classical recrys-
Dislocation pipe diffusion [41] 0.18 20.34
Bulk/lattice diffusion [40] 0.035 28.8
tallization [42,43] plus strain induced boundary migration (SIBM)
[43–46]. SIBM was shown to leave orientation gradients [44]. This
appears to be the only rationale for clear orientation gradients in
grains after ‘apparent’ fully recrystallization (Fig. 9c). s13 stress re-
lief remained incomplete, see Fig. 5. Prolonged annealing, and vir-
brought down such differences. During the final cold rolling tual elimination of orientation gradients, brought the stress relief
cladding and surrogate meat had similar strains. In other words, to completion.
initial unconstrained deformation of the picture-frame assembly This study provides an important correlation between near
latter became constrained. A constrained deformation, with dif- interface out of plane residual stress (s13) and bond strength. This
ferences in elastic–plastic properties, is expected to develop is of practical relevance. More importantly, however, the present
residual stresses [42]. It was stipulated that constrained defor- study brings in important phenomenological correlations between
mation of cladding and surrogate meat resulted in ‘lateral’ stres- microstructure and stress relief. If stress is considered as a ‘distor-
ses (Fig. 9a(ii)). In extreme situation, generation of near interface tion’ of the unit cell (changes in inter planar spacings and angles)
microscopic shear bandings was also hypothesized (Fig. 9a(iii)). [42], then such a distortion needs to be supported by appropriate
The presence of near interface s13 component (Fig. 2d) and features of deformed microstructures. Similarly, stress relief is a
microscopic shear banding (Fig. 9b) supports the simplified sche- ‘practical’ topic: which also should have microstructural origin.
matics/hypothesis of Fig. 9a. It is apparent that developments in The present study supports these arguments and perhaps initiates
strong s13 also accompanied formation of the shear bands. future discourse on the ‘microstructural origin of stress relief
Annealing at 300 °C selectively eliminated such shear bands. This annealing’.
158 D. Kohli et al. / Journal of Nuclear Materials 447 (2014) 150–159

(a)

(b)

(c)
Fig. 9. (a) Hypothesizing developments in lateral stresses and shear banding during fabrication of simulated plate fuel. Details included in the text. (b) Direct EBSD
observations and corresponding image quality maps showing elimination of shear bands at 300 °C (30 min) stress relief. (c) ‘Apparent’ full recrystallized (500 °C and 30 min
annealing) grains had orientation gradients. Such gradients are shown with respect to average grain orientations in appropriate color codes’’.

5. Conclusions Acknowledgements

An inverse relation between measured bond strength and esti- The authors would like to acknowledge the EBSD and stress
mated near interface out of plane residual shear stress (s13) was measurements from the National Facility (DST-IRPHA) at IIT Bom-
established during the final stages of simulated plate fuel element bay. Support from BRNS (board of research in nuclear science) is
fabrication. The relationship was observed after the final fabrica- greatly appreciated. Support provided by the team members at
tion (cold rolling and straightening), as well during stress relief Plate Fuel Development and Fabrication group at MFD (metallic
annealing. The latter enabled drop in s13 with corresponding in- fuel division, BARC) was critical: special acknowledgments to Mr.
crease in bond strength. M.R. Shaikh, S.S. Gotad, B.K. Mondal, B.N. Pisal and R.M. Malagi.
Developments in ‘lateral’ stresses (s13) were supported by near
interface microscopic shear bands. Annealing at 300 °C eliminated References
such bands through partial recrystallization and recovery. This was
more effective than stress relief at higher (400 and 500 °C) temper- [1] F. Thummler, H.E. Lilienthal, S. Nazare, Powder Metall. 12 (1969) 1–22.
[2] James A. Lake, Donald K. Parsons, John L. Liebenthal, John M. Ryskamp, Gary N.
atures. The latter formed large grains, possibly through combina-
Fillmore, Deslonde R. Deboisblanc, Nucl. Instrum. Methods Phys. Res. A249
tions of classical recrystallization and SIBM. These grains, (1986) 41–52.
however, had orientation gradients: and incomplete relief in s13. [3] M.K. Meyer, T.C. Wiencek, S.L. Hayes, G.L. Hofman, J. Nucl. Mater. 278 (2000)
Prolonged annealing plus elimination of orientation gradients were 358–363.
[4] Dennis D. Keiser Jr., Adam B. Robinson, Jan-Fong Jue, Pavel Medvedev, Daniel
needed to accomplish full stress relief. M. Wachs, M. Ross Finlay, J. Nucl. Mater. 393 (2009) 311–320.
D. Kohli et al. / Journal of Nuclear Materials 447 (2014) 150–159 159

[5] M.K. Meyer, G.L. Hofman, S.L. Hayes, C.R. Clark, T.C. Wiencek, J.L. Snelgrove, [23] P. Van Houtte, L. De Buyser, Acta Metall. Mater. 41 (2) (1993) 323–336.
R.V. Strain, K.H. Kim, J. Nucl. Mater. 304 (2002) 221–236. [24] I.C. Noyan, J.B. Cohen, Residual Stress-Measurement by Diffraction and
[6] Dennis D. Keiser Jr., J. Gana, J.F. Juea, B.D. Miller, C.R. Clark, Mater. Interpretation, Springer-Verlag, Berlin, 1987.
Characterization 61 (2010) 1157–1166. [25] M.E. Fitzpatrick, A.T. Fry, P. Holdway, F.A. Kandil, J. Shackleton, L. Suominen
[7] G.L. Hofman, L.A. Neimark, F.L. Olquin, 1986 International Meeting on Reduced Determination of Residual Stresses by X-ray Diffraction-Issue 2, DTI
Enrichment for Research and Test Reactor, U.S.A., November 3–6, 1986. Measurement Good Practice Guide No. 52, 2005.
[8] C.R. Clark, S.L. Hayes, M.K. Meyer, G.L. Hofman, J.L. Snelgrove, Transactions of [26] A. Baczmanski, C. Braham, W. Seiler, N. Shiraki, Surf. Coat. Technol. 182 (2004)
session 2, in: 8th International Topical Meeting on Research reactor and fuel 43–64.
management, March 21–24, 2004. [27] P. Gergaud, S. Labat, O. Thomas, Thin Solid Films 319 (1998) 9–15.
[9] Hakan Ozaltun, J. Nucl. Mater. 419 (2011) 76–84. [28] J. Peng, V. Ji, W. Seiler, A. Tomescu, A. Levesque, A. Bouteville, Surf. Coat.
[10] Gregory K. Miller, Douglas E. Burkes, Daniel M. Wachs, Mater. Des. 31 (2010) Technol. 200 (2006) 2738–2743.
3234–3243. [29] B.D. Cullity, S.R. Stock, Elements of X-ray Diffraction, third ed., Prentice Hall,
[11] Hakan Ozaltun, Herman Shen and Pavel Medvedev, Proceedings of the ASME 2001.
2011 International Mechanical Engineering Congress & Exposition, November [30] T. Unger, A. Borbely, Appl. Phys. Lett. 69 (1996) 3173–3175.
11–17, 2011, Denver, Colorado, USA. [31] I. Groma, Phys. Rev. B 57 (1998) 7535–7542.
[12] C. Liu, M.L. Lovato, D.J. Alexander, K.D. Clarke, N.A. Mara, W.M. Mook, M.B. [32] G. Guiglionda, A. Borbely, J.H. Driver, Acta Mater. 52 (2004) 3413–3423.
Pime, D.W. Brown, in: 34th International Meeting on Reduced Enrichment for [33] R. Khatirkar et al., ISIJ Int. 51 (2011) 849–856.
Research and Test Reactors, October 14–17, 2012. [34] A. Albou, S. Raveendra, P. Karajagikar, I. Samajdar, C. Maurice, J.H. Driver,
[13] R. Rakesh, D. Kohli, I. Samajdar, V.P. Sinha, G.J. Prasad, Fabrication of simulated Scripta Mater. 62 (2010) 469–472.
plate fuel elements:defining role of out-of-plane residual shear stress, J. Nucl. [35] Jérémy Epp, Holger Surm, Thomas Hirsch, Franz Hoffmann, J. Mater. Process.
Mater 445 (2014) 200–208. Technol. 211 (2011) 637–643.
[14] S.H. Lee, Y. Saito, T. Sakai, H. Utsunomiya, Mater. Sci. Eng. A325 (2002) 228– [36] O. Vohringer, Advances in surface treatments, Intr. Guidebook on Residual
235. Stress, vol. 4, Pergamon Press, Oxford, 1987. pp. 367–396.
[15] S.H. Lee, Y. Saito, N. Tsuji, H. Utsunomiya, T. Sakai, Scripta Mater. 46 (2002) [37] Weizhi Luan, Chuanhai Jiang, Vincent Ji, Mater. Trans. 50 (2009) 1499–1501.
281–285. [38] D.A. Porter, K.E. Easterling, Phase Transformations in Metals and alloys, second
[16] K.H. Eckelmeyer, Advances in Surface Treatment, vol. 4, Pergamon Press, 1985. ed., Chapman & Hall, UK, 1992.
[17] M. Movahedi, H.R. Madaah-Hosseini, A.H. Kokabi, Mater. Sci. Eng. A 487 (2008) [39] P.G. Shewmon, Diffusion in Solids, second ed., McGraw-Hill, New York, 1963.
417–423. [40] M.F. Ashby, Acta Metall. 20 (1972) 887–897.
[18] T. Ericsson, The effect of final shaping prior to heat treatment, in: G. Totten, M. [41] Jurgen Mimkes, Thin Solid Films 25 (1975) 221–230.
Howes, T. Inoue (Eds.), Handbook of Residual Stress and Deformation of Steel, [42] B. Verlinden, J. Driver, I. Samajdar, R.D. Doherty, in: R.W. Cahn (Ed.), Thermo-
ASM International, 2002, pp. 150–158. Mechanical Processing of Metallic Materials, ISBN-978-0-08-044497-0,
[19] ASTM: D 1876-01, Standard Test Method for Peel Resistance of Adhesives (T- Pergamon Materials Series – Series, Elsevier, Amsterdam, 2007.
Peel Test). [43] F.J. Humphreys, M. Hatherly, Recrystallization and Related Annealing
[20] ASTM: C 633-01, Standard Test Method for Adhesion or Cohesion Strength of Phenomena, second ed., Pergamon Press, Oxford, 2004.
Thermal Spray Coatings. [44] S. Raveendra, H. Paranjape, S. Mishra, H. Weiland, R.D. Doherty, I. Samajdar,
[21] J.S. Cheon, Y.S. Kim, Material Properties of Aluminum Alloys and Pure Met. Trans. A 40A (2009) 2220–2230.
Zirconium for Use in High-density Fuel Development for Research Reactors, [45] Paul A. Beck, Philip R. Sperry, J. Appl. Phys. 21 (1950) 150–152.
ANL/RERTR/TM-12-6, 2012. [46] Peter Bate, Bevis Hutchinson, Scripta Mater. 36 (1997) 195–198.
[22] U. Welzel, J. Ligot, P. Lamparter, A.C. Vermeulen, E.J. Mittemeijer, J. Appl. Cryst.
38 (2005) 1–29.

You might also like