Sie sind auf Seite 1von 11

Nuclear Engineering and Design 320 (2017) 298–308

Contents lists available at ScienceDirect

Nuclear Engineering and Design


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

Sub-size tensile specimen design for in-reactor irradiation and


post-irradiation testing
Maxim N. Gussev ⇑, Richard H. Howard, Kurt A. Terrani, Kevin G. Field
Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA

h i g h l i g h t s

 New specimen geometry for post-irradiation testing (PIE) was designed.


 The new geometry performance was evaluated for several commercial and model alloys.
 Scale factor role was discussed in detail.

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

Article history: The present work aims to provide a complete engineering solution, an appropriate experimental data-
Received 12 October 2016 base, and a brief physical background on designing a miniature specimen geometry suitable for irradia-
Received in revised form 30 May 2017 tion in materials test reactors such as the High Flux Isotope Reactor and post-irradiation out-of-hot cell
Accepted 3 June 2017
testing. The physical limits of specimen miniaturization and a background of the scale factor effect are
discussed, and principal limitations are defined. The advantages of modern test methods like digital
image correlation, as well as some limitations connected to small specimen size, are analyzed. A sub-
Keywords:
sized specimen geometry, ‘‘SS-Mini,” is designed; the geometry employs existing irradiation capsules
Scale factor
Sub-size specimen geometry
leading to the reduced cost of any irradiation campaign. The new geometry performance is evaluated
Post-irradiation investigation using a commercial 304 L stainless steel, an aluminum alloy including advanced 3D-printed material, a
In-situ testing high nickel 718-alloy, tungsten, and an advanced fuel cladding FeCrAl alloy. Mechanical tests are con-
Flexible design for neutron irradiation ducted to compare the engineering mechanical properties (yield and ultimate tensile stress, uniform
capsules and total elongation values) and plastic behavior of the proposed miniature specimen with common
specimen types for irradiation testing.
Ó 2017 Elsevier B.V. All rights reserved.

1. Introduction mechanical properties or at a minimum yield coherent scaling


parameters from small scale testing. The motivation for small-
There is a constant and growing interest in the mechanical test- scale mechanical testing in the nuclear industry is to best take
ing of small specimens and analyzing material behavior at small advantage of the limited experimental space in materials test reac-
scales. In general materials science, this interest is driven by intro- tors and minimize the radiological dose produced from activated
ducing smaller devices and exploring the micro- and nano- materials that directly scales with their volume.
mechanics areas (Connolley et al., 2005; Jaya and Alam, 2013). Additionally, recent progress in different in-situ test methods,
The goal is to analyze the properties and performance of micro- where small scale testing is routinely needed, is impressive
parts and components in detail while understanding materials (Dehm et al., 2006), and the next decade is likely to see a consider-
behavior at lower length scales. able rise in the in-situ test techniques for general materials science
At the same time, the nuclear industry often has a different goal and nuclear engineering. For instance, micro-pillar testing (Dehm
(Klueh, 1985; Kohno et al., 2000; Hindley et al., 2015; Viehrig et al., et al., 2006; Shin et al., 2015), high-resolution EBSD (Jiang et al.,
2015): testing of miniature samples taken from nuclear installa- 2013), advanced tools for 3D structure reconstruction, and other
tions should provide information that is applicable to bulk techniques are quickly providing new, unique, and—most impor-
tantly—insightful data. Employing these methods is an attractive
idea for the investigation of radiation effects in materials, including
⇑ Corresponding author at: PO Box 2008, Oak Ridge, TN 37831, USA. plastic strain at small scales (Greer and De Hosson, 2011).
E-mail address: gussevmn@ornl.gov (M.N. Gussev).

http://dx.doi.org/10.1016/j.nucengdes.2017.06.008
0029-5493/Ó 2017 Elsevier B.V. All rights reserved.
M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308 299

Moreover, the development of innovative nuclear systems con- samples are of interest. Reaching the acceptable level of radioactiv-
stantly requires new materials, which, in turn, means increasing ity may require unacceptable waiting times that exceed the life-
needs for advanced comprehensive testing methods (Fazio et al., time of the funding program or project interested in such data.
2011; Yvon et al., 2015), like in-situ corrosion testing (Klecka Thus, there is a need for a specimen geometry capable of sup-
et al., 2015), while simultaneously reducing the time to achieve porting the modern in-situ testing methods on moderate to highly
such data and analysis. neutron irradiated materials, while still providing repeatable bulk-
The deployment of the advanced in-situ techniques on neutron scale mechanical properties. It is important for the new geometry
irradiated materials is of special importance. Although significant to provide appropriate mechanical test results and keep a continu-
progress was made to use light ion and heavy ion irradiations to ity and comparability with the existing databases in the literature.
simulate neutron irradiation while producing samples with little If suitable, these small-scale mechanical testing specimens impact
or no radioactivity [e.g., see (Was et al., 2002)], the discrepancies scientific productivity in the nuclear materials area in two impor-
in the observed microstructure and mechanical properties still tant ways. First, by avoiding in-cell testing, a whole host of new
exist. Hence, data for neutron irradiated materials are still essential characterization techniques too costly or simply inviable for in-
to fully develop a nuclear-grade alloy ripe for commercial cell deployment—where high radiation fields detrimentally inter-
deployment. act with electronic components—suddenly become available for
Irradiation may stimulate specific processes not observed with- utilization. Second, the costs associated with in-cell testing are
out irradiation, including swelling, increased diffusion and creep, often very high (e.g., at least a factor of 2–3x higher at ORNL)
radiation-induced segregation (RIS), and irradiation-assisted stress and would, in turn, limit the depth and breadth of the R&D project.
corrosion cracking (IASCC), just to name a few. Advanced in-situ The circumstances discussed above motivated the development
tools may bring new insights into well-known processes. For of a new specimen geometry allowing for: (1) in-HFIR irradiation
instance, a preliminary, not-on-purpose analysis of the deforma- within the existing ‘‘rabbit capsule” geometry and specifications,
tion localization in neutron irradiated austenitic steels revealed (2) out-of-hot cell testing with short ‘‘cool-down” (reduction in
specific phase transformation at the irradiated specimen’s surface radioactivity by decay) time, and (3) the capability to perform in-
(Gussev et al., 2014b) along with areas of localized deformation situ measurements using advanced techniques. The present work
and high dislocation density spots (Field et al. 2014a) in addition discusses the design, geometry, and implementation of a miniature
to the expected and well-known defect-free channels. It is impor- sub-size specimen that meets these requirements.
tant to be able to test the irradiated specimen in-situ, registering,
for example, the evolution of dislocation density and the evolution
of local misorientation (Jiang et al., 2013) with direct measure- 2. The background of scale and size effects
ments of acting stress or observations of stress corrosion crack
appearance and propagation. These data, coupled with modern The scale factor and size effects usually become apparent when
modeling tools, allow for complex analysis of material behavior, the specimen size approaches the characteristic length scale of the
including constitutive crystal plasticity approaches. However, to material microstructure (Connolley et al., 2005). For common
reach these goals, one has to at least have direct access to the spec- steels and alloys, this tends to be the grain size. Another size effect,
imen for preparing the appropriate quality surface. often observed in brittle materials (Bažant, 1999), occurs when the
Miniature sub-size specimens for post-radiation testing have a size of the specimen becomes small compared to the spatial distri-
long history and are widely used for investigating mechanical bution and density of critical defects. Additionally, size effects may
properties and deformation-hardening behavior of irradiated and appear due to physical, chemical, or corrosion processes that are
non-irradiated metals and alloys (Klueh, 1985; Panayotou et al., negligible for relatively large objects but become critical when
1986; Kohyama et al., 1991; Kohno et al., 2000; Gussev et al., the size of the test specimen decreases. Most of these processes,
2014a), weldments (Field et al., 2014b), nanostructured materials, like oxidation, usually do not appear in common mechanical tests;
and different composites. A variety of different sizes and geome- however, some important physical phenomena sensitive to the
tries are used in nuclear materials science (Klueh, 1985; Pierron scale factor are briefly analyzed below.
et al., 2003; Wakai et al., 2011), with new geometries and testing Considering the scale factor role, researchers most often focus
methods being constantly developed (Rickerby and Fenici, 1986; on the uniform strain area, where the uniaxial stress state exists,
Džugan et al., 2014; Hurst and Matocha, 2015; Rund et al., 2015). and on the necking behavior, where the complex stress state and
Specimen miniaturization often involved other methods, not only the stress triaxiality play an important role.
tensile testing. Thus, Jung et al., (1996) offered a miniature tensile In the first case, under uniaxial stress, the specimen is often pre-
specimen with gauge section of 5  1  0.4 mm; additionally, a sented as a ‘‘multilayer” object with a layer of near-surface grains.
number of testing methods for miniature specimens (tensile, fac- The ratio of the ‘‘near-surface grains” to the ‘‘bulk grains” impacts
ture toughness, punching, impact testing) were discussed in detail the specimen deformation behavior and strain hardening rate
(Jung et al., 1996) focusing on the detailed correlation between (Wang et al., 2013; Shin et al., 2015). The near-surface grains (or
miniature and bulk behavior. The offered tensile specimen geome- the near-surface layer if a single crystal specimen is considered)
try was intensively used in a number of projects (see for instance, have smaller dislocation density and internal back stress level
(Dai and Bauer, 2001)). compared to the bulk grains (Keller et al., 2010; Shin et al., 2015)
Through the past few decades, there has been a tendency to because of the dislocation escape through the free surface. Also,
decrease the overall specimen size and volume (Klueh, 1985). Prior the surface plays a key role in crack initiation during fatigue
to this work, two miniature tensile specimen geometries were (Signor et al., 2016) and stress corrosion cracking (SCC). To handle
commonly used for irradiations in the High Flux Isotope Reactor very thin specimens (10–20 lm sheets), some special test method
(HFIR): the SS-3 and the SS-J type specimens (Klueh, 1985; (s) may be needed. For instance, Hoffman and Hong (Hoffmann and
Gussev et al., 2014a). These two geometries provide acceptable Hong, 2006) offered an aero-bulging test method for testing thin
mechanical properties compared to standard-scale specimens foils of different materials.
(Gussev et al., 2014a). However, the post-irradiation activity level It is worth noting that the near-surface layer is not necessarily
is usually too high for nuclear reactor metals after discharge, and ‘‘weak;” that is to say, it does not necessarily have a decreased
this limits the post-irradiation analysis to testing within a hot cell strength. Often, specimens are produced by electric discharge
facility only, especially when moderate to high damage dose (dpa) machining (EDM) followed by mechanical grinding or polishing
300 M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308

to fit the dimensional requirements. EDM leads to the formation of However, the total elongation value demonstrated a more com-
a recast surface layer, often referred to as the ‘‘white layer” (Bleys plex behavior, as shown in Fig. 1, where saturation in the total
et al., 2006), which may be hard and brittle depending on the elongation values was reached at 2 mm thickness (T/W ratio of
material system. The subsequent mechanical grinding removes 0.67 in Fig. 1). Necking angle, the angle between the loading axis
the EDM layer but introduces cold-work related artifacts or even and the fracture plane, was also affected by thickness. The param-
results in the formation of fine grains (Kaneda et al., 2011). Elec- eter controlling these processes was found to be the T/W ratio that
tropolishing or controlled chemical etching may eliminate the influences the stress state in the neck (Byun et al., 1998). A similar
modified layer, but these techniques are both costly and time- result—post-necking elongation increased with thickness—was
consuming or produce specimens with thickness and edge shape obtained by (Yuan et al., 2012). In the context of the present work,
variations that can significantly skew results. this means that the T/W value of the conventional small-scale
Strictly speaking, it is impossible to avoid the free surface influ- specimens, like SS-3 [T/W  0.5, (Klueh, 1985)], and new miniature
ence. Instead, the question is how thick should the specimen be to geometries discussed later should be close, if not identical.
provide acceptable data? Usually, in mechanical test practice, the If the material behavior under the complex stress state is of
smallest specimen dimension should exceed 5–6 grains to mini- interest, a proper specimen geometry may be employed to analyze
mize the scale factor influence (Kohno et al., 2000). This rule may the stress triaxiality role on plastic behavior and fracture, as dis-
not be fully valid in some specific cases (e.g., in nano-structured cussed recently in (Algarni et al., 2015). If necessary, the stress
materials with high defect density). If defect spacing is much smal- state may be directly addressed via neck geometry measurements
ler than grain size, the bulk mechanical properties may be mea- [e.g., by multi-camera 3D DIC, (Kamaya and Kawakubo, 2014)] or
sured by using smaller specimens that contain as little as 2–4 by a finite element modeling (FEA) approach (Kim and Byun, 2010).
grains across the smallest dimension (Howard et al., 2016). A number of additional phenomena exist beyond the few dis-
Sometimes, there are additional limitations if the material being cussed above (Jaya and Alam, 2013); however, a detailed analysis
tested exhibits phenomena sensitive to the active volume or sur- is out of the present scope of work. In summary, it appears that
face conditions. For instance, plastic deformation generates energy geometrical (specimen thickness and presence of natural internal
release in the form of heat and leads to a temperature increase boundaries, like grain boundaries), technological (surface and
(Ayres, 1985), but this process is usually ignored. However, the near-surface layer conditions), and physical (potential scale-
temperature at the specimen gauge may rise by 10 °C or more even sensitive phenomena) limitations should be analyzed for the par-
at a moderate strain rate (10 3 s 1) (Ayres, 1985). In the case of ticular material before designing a specimen geometry for nuclear
fracture, a large amount of energy is being released in a small vol- materials development applications.
ume resulting in a temperature jump that can reach 100 °C. For
example, a temperature rise of 300 °C was observed in the shear
3. Materials and test methods
bands in metallic glass specimens (Yang et al., 2006). As an ulti-
mate case, melting of a metallic glass specimen during fracture
As discussed above, the scale-factor limiting criteria can vary
was observed in-situ (Yang et al., 2004). In the tensile test, the heat
from material to material and can be significant or insignificant
generation is proportional to the specimen volume (d3) in the
depending on the data of interest from a mechanical test. Thus, it
uniform strain area, whereas the heat losses rely on the specimen
is important to cover a wide range of strength and ductility values
surface or cross-sectional area (d2). Thus, the heating intensity is
and different types of strain hardening behavior by using a vast
proportional to the specimen size (d), being less pronounced in
array of different material systems. The present work included
small specimens compared to larger ones. If the material of interest
both well-known and advanced modern materials, as shown in
experiences strain-induced phase instability (Chen et al., 2013;
Table 1. Aluminum 6061 alloy is employed to cover the low
Shirdel et al., 2015), such as the transformation induced plasticity
strength/low ductility range; also, after annealing, this material
(TRIP) effect, deformation twinning, or dynamic deformation aging,
demonstrates specific deformation behavior with propagating
the effects of strain-induced heating should not be ignored.
deformation bands. Austenitic AISI 304L stainless steel is selected
Additionally, some materials, like carbon steels and Al-Mg
as a representative commercial, widely used nuclear material.
alloys, often exhibit Luders band propagation. In this case, a defor-
mation plateau is observed in the tensile curves (Mazière and
Forest, 2015). It appears that Luders deformation may also be
scale-sensitive. For instance, the plateau disappeared in a ferritic
A533B steel when a thin specimen geometry such as an SS-2 type
tensile specimen (T = 0.25 mm) is used (Gussev et al., 2014a). Also,
the material of interest may be sensitive to texturing; orientation
effects may appear in the specimens produced from rolled or
extruded material (Zhang et al., 2014).
The scale factor’s role during deformation localization and neck-
ing may be more complex compared to the uniform strain area. In
addition to the free surface effects and the different behavior of the
near-surface grains, the neck also experiences a complex stress
state. The neck shape and principal strain ratio may depend on
specimen thickness and thickness-to-width ratio (T/W). Byun
et al. (Byun et al., 1998) investigated the effect of specimen thick-
ness on the mechanical properties of miniature specimens with a
particular focus on deformation localization and necking. Regard-
ing the yield and ultimate stress and uniform ductility for ferritic
Fig. 1. The effect of the T/W ratio on the uniform (UE) and total (TE) elongation
steels, Byun et al. concluded (Byun et al., 1998) that these values
values. The plot includes the data from (Byun et al., 1998) on an RPV steel and the
will be close to the bulk material data of the specimen if thickness present work’s results on the advanced FeCrAl alloy (see below). The positions of
exceeds  0.3–0.4 mm (or  7–10  grain size). the designed small specimens on the T/W scale are shown by vertical, dashed lines.
M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308 301

Table 1 widely discussed in the literature, see (Sutton et al., 2009) and
Element composition of the investigated materials. those cited within.
Material Composition, wt% (or ppm wt. for tungsten).
6061 Al- Al: bal., Mg: 0.86, Si: 0.66, Fe: 0.35, Cu: 0.23, Cr: 0.13, Ti: 0.08,
alloy Mn: 0.08, Zn: 0.08, Ni: 0.02 4. Miniature specimen design and properties
304 L Fe: bal., C: 0.02, Mn: 1.3, Si: 0.45, Cr: 18.3, Ni: 8.02, Mo: 0.074
FeCrAl- Fe: 79.4, Cr: 13.1, Al: 5.3, Y: 0.05, Mo: 2.0, Si: 0.13, Nb: <0.01
4.1. Design and geometry considerations
alloy
718-alloy Ni: bal., Cr: 18.5, Fe: 18.7, Mn: 0.23, Mo: 3.03, Al: 0.48, Co: 0.29.
Tungsten (ppm wt.) W: bal., Al: 0.23, Si: 0.55, P: 0.68, V: 0.82, Cr: 1.7, Fe: 3, The following aspects based on the prior discussion were con-
Ni: 0.75, As: 0.39, Sn: 0.69, Ta < 5, other: <0.1 ppm wt. sidered during the sub-size specimen design phase:

1. The miniature specimen geometry should reproduce a


Commercial 718-alloy is added as a high strength material. polycrystalline-like behavior. To achieve this, the specimen
Advanced FeCrAl alloy is selected to represent a system currently should contain at least 5–6 (or more) grains per the smallest
undergoing alloy development efforts for accident tolerant clad- dimension. Generally, if the thinnest area has less than 5 grains,
ding and hence a material with a limited material database (Field the specimen will demonstrate reduced strength. The most
et al., 2014b; Yamamoto et al., 2015). Tungsten was selected as a common grain size for ferrous and non-ferrous alloys is near
material to exhibit high strength, but low, if not zero, ductility. 40–50 lm, resulting in 300–400 lm being the limiting mini-
To provide a baseline for comparison, Fig. 2 shows the geometry mal thickness for the specimen.
and dimensions of the common specimen (the SS-J type with 2. To accurately reproduce necking and strain localization behav-
0.75 mm thickness) used in the present work. The SS-J type sample ior (or, at an absolute minimum, close to the SS-J types), the
shown in Fig. 1 is typically used as the sheet type dog bone speci- geometry should meet specific thickness/width and length/
men. This geometry provides effective heat transfer across the cross-section requirements.
stacked faces during irradiation. Shoulder loading without pin- 3. The specimen should have limited activity after irradiation.
holes provides sufficient material in the head of the tensile speci- Currently, at the ORNL’s LAMDA (Low-Activation Materials
men for microhardness testing and electron microscopy sample Development and Analysis) (Parish et al., 2015) facility, the
preparation. limit is 100 mR/h at 30 cm per specimen to allow for the out-
Both SS-J and SS-Mini specimens discussed below are produced of-hot cell research activity. This limit is generally within typi-
from the bulk material using EDM. The specimens are also ground cal limits for many research facilities within the United States.
down mechanically by a vendor to fit the dimensions. All speci- Thus, specimen volume should be significantly reduced, com-
mens are produced by the same vendor. Before the tensile test, pared to an SS-J, especially if high dose (>10 dpa) specimens
the thickness and width of each specimen are measured with an are of interest.
accuracy of 10 lm or better; gauge length variations (±0.02 mm) 4. Sample geometry should be simple and avoid or minimize
are too small to influence the plasticity results. For each geometry rounded shapes to reduce the sample production cost. In the
and material, 4–5 specimens are tested per point. The strain rate is future, it would be advantageous to allow for machining the
close to 10 3 s 1 (with no more than a 5% difference) for all spec- same samples from irradiated materials in a hot cell facility.
imen geometries. 5. Additionally, the specimens should fit into existing irradiation
Tensile tests are performed on an MTS Insight 2–52 one-column capsule designs, thereby eliminating the need for conducting
tensile screw machine. All tensile specimens are shoulder loaded complex heat transfer calculations. The number of specimens
and tested at room temperature. Before the tensile tests, a subset per capsule should be larger compared to the SS-J types, allow-
of specimens are painted with a random speckle pattern. High- ing for larger statistics or distributing the specimens among the
resolution Allied Vision GT6600 and GX3300 cameras are used in collaborating groups. Additionally, each capsule should be able
the experiments; the lenses used provide a resolution of 10– to carry both new and old specimen types and optimize the
15 lm per pixel for the GT6600 and 2–2.5 lm per pixel for the number of specimens of each type as necessary for a particular
GX3300 for both the SS-J type and miniature specimens. Strain project.
fields and true stress–true strain curves are calculated using VIC-
2D commercial software and a custom program utilizing common
digital image correlation (DIC) algorithms. DIC is a modern, con- 4.2. The geometries included in the preliminary analysis
stantly developing experimental tool (Sutton et al., 2009) that
allows for non-contact strain measurements; the details, peculiar- After analyzing the criteria and limitations listed above, several
ities, and limitations of the DIC-based strain measurements are geometries were considered for the preliminary analysis, as shown
in Fig. 3. Two gauge lengths were chosen: 2.55 mm and 3.55 mm
(Types 1 and 2, respectively), and two head types were employed.
Small heads (S) were expected and allowed to deform during the
test; however, these heads were strong enough to carry the load.
For this geometry, the head shape stability was sacrificed to reduce
mass and radioactivity level after neutron irradiation. As will be
discussed below, DIC allows for retrieving correct true stress–true
strain curves for the gauge portion of the specimen, regardless of
the selected head geometry. Extended heads (E) were designed
to carry the load with minimal plastic strain in the head. Further
head size increase would eliminate the plastic strain in the heads
Fig. 2. Geometry and dimensions of the common miniature specimen (SS-J type). completely, but the price paid would be the extra sample mass
Dimensions are in mm; ‘‘T” designates thickness. Dimensional tolerances were as
follows: gauge length ±0.02 mm, gauge width ±0.02 mm, specimen thickness
and extra radioactivity after neutron irradiation. For each small
±0.01 mm, and other dimensions ±0.04 mm. Grips used to test the specimen are specimen geometry (Types 1 and 2, S- and E-heads), two different
shown in (Gussev et al., 2014a). specimen thicknesses (0.4 mm and 0.6 mm) were considered. All
302 M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308

Fig. 3. General view of the preliminary geometries (left) and the detailed geometry of the Type-2E (right).

Table 2
A number of specimens per irradiation capsule and mechanical properties of the investigated FeCrAl alloy.

Mechanical properties
Specimen geometry Specimens per standard Yield stress, Ultimate Uniform Total Grading
HFIR rabbit capsule MPa stress, MPa elongation,% elongation,% points
SS-J type 12  2* 710 ± 25** 767 ± 28 7.9 ± 2.5 14.8 ± 3.7
SS-Mini*** Short 1S-0.4 12  9 607 760 10.8 23.1 
(2.55 mm gauge)
1E-0.4 12  9 725 782 6.6 18.2 
1S-0.6 12  6 663 796 10 29 –
1E-0.6 12  6 761 826 9 28 
Long 2S-0.4 12  6 708 821 9 18 
(3.55 mm gauge)
2E-0.4 12  6 705 ± 50 761 ± 52 7±1 15 ± 2 
2S-0.6 12  4 702 822 11 25 
2E-0.6 12  4 753 821 9 22 
*
For specimen with 0.75-mm thickness, 123 if an SS-J (T = 0.5 mm) type specimen is used.
**
One standard deviation value is given for two geometries as an inaccuracy estimation.
***
Nomenclature example: the ‘‘1S-0.4” abbreviator means Type 1 (short gauge) with small (S) head and thickness of 0.4 mm.

miniature specimens, regardless of thickness and gauge length, specimen requires fixed mechanical grips and has higher mass
could be tested using the same grips. and volume. In the present work, pneumatic grips were rejected
Table 2 shows the number of different specimens allowable per since they require much larger heads to fix the specimen in a
capsule using standard packing configurations for irradiation cap- timely and reliable manner—a key factor when working with
sules in the central flux position of the HFIR. As discussed later highly radioactive samples. The grips geometry employed within
in Section 5, each irradiation capsule (commonly referred to as this work (Gussev et al., 2014a) provides some degrees of freedom
‘‘rabbits”) contains three small cylindrical holders with four stack to allow for alignment of the specimen to the loading axis.
positions each. Thus, there may be 12 stacks total, with two SS-J
type 0.75 mm thickness specimens in each—or 24 SS-J type [Thick- 4.3. Geometry comparison and selection
ness (T)=0.75 mm] specimens in total. Using miniature geometries,
each stack may contain a larger number of small specimens, from 9 The mechanical test results obtained for the SS-Mini geome-
(SS-Mini 1E-0.4) to 4 (SS-Mini 2E-0.6). Specimens of different types tries are compared with the data obtained for the SS-J type geom-
may be combined without disturbing the capsule’s internal tem- etry. Table 2 demonstrates a typical dataset for the FeCrAl alloy.
perature field and dose distribution; this aspect is discussed in Scattering (one standard deviation value) is shown for several
greater detail in Section 5. geometries.
Comparable specimens are constantly being developed and are Here, close matching (i.e., the known, repeatable deviation) in
used by a number of authors. For instance, small specimen the mechanical properties was deemed beneficial as it allows for
geometry with a gauge of 2 mm  1 mm  0.2 mm was designed, comparing the new data with an existing database of mechanical
irradiated, and used to investigate the radiation tolerance of properties. Thus, using the data in Table 2, the best geometry
ultra-fine grained steel (Alsabbagh et al., 2013). Additionally, was chosen among the analyzed set using the following logic:
Džugan et al. (2014) offered a tensile specimen with a
3 mm  1.5 mm  0.5 mm gauge; the specimen may be produced 1. The specimens with small heads (S) demonstrated, as a rule,
from an 8 mm diameter disk. It was shown that non-contact optic smaller yield stress and larger ductility (especially total elonga-
measurements (e.g., DIC) may provide additional important infor- tion) compared to the SS-J type. The extended heads (E) pro-
mation and true stress–true strain curves; the offered geometry vided much better matching of the yield stress with less than
(Džugan et al., 2014) provides good tensile test results in compar- 10% difference compared to the SS-J type, Table 2. To grade
ison with standard geometries for some common materials the different geometries, all E-headed specimens received one
(aluminum alloy, 99.99% copper, 14% chromium–steel, etc.). point in the selected ranking system for the selection of the best
However, compared to the geometries proposed here, their geometry.
M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308 303

Table 3
Mechanical properties of the tested materials (SS-J vs. SS-Mini 2E-0.4).

Material Geometry Yield stress, MPa Ultimate stress, MPa Uniform elongation,% Total elongation,%
SS-Mini feasible materials
6061-0 SS-J 72 122 18.1 27.3
SS-2E 73 119 16.9 26.5
6061-T6 SS-J 294 315 7.2 15.4
SS-2E 279 301 5.8 14.3
Annealed 304 L SS-J 274 ± 22* 789 ± 14 67 ± 3 76 ± 3
SS-2E 272 ± 23 781 ± 34 69 ± 3 80 ± 4
FeCrAl SS-J 710 767 7.9 14.8
SS-2E 705 761 7.0 15.0
718-alloy SS-J 1237 ± 13 1461 ± 25 16.3 ± 0.4 22.5 ± 0.9
SS-2E 1170 ± 33 1386 ± 36 15.4 ± 0.8 21.0 ± 1.5
Non-SS-mini feasible
Tungsten, W SS-J 820** – – –
SS-2E 603** – – –
*
One standard deviation is shown for several materials to illustrate the typical result scattering.
**
Fracture stress is given for tungsten.

2. The specimens with a short gauge (Type 1, 2.55 mm) demon-


strated slightly higher uniform elongation and higher total
elongation compared to the specimens with a long gauge (Type
2, 3.55 mm); this result agrees with (Pierron et al., 2003). Duc-
tility values for the Type 2, in general, were closer to the SS-J
type compared to the Type 1. This provided an additional selec-
tion point for the Type 2 specimen group.
3. The thick specimens (0.6 mm) always had larger total elonga-
tion compared to the thin ones (0.4 mm). Thin 0.4 mm speci-
mens were closer to the SS-J type, thus providing the third
selection point as shown in Table 3. The resulting grading indi-
cated the SS-Mini 2E-0.4 geometry (or shortened, the SS-2E,
Fig. 3) as the best SS-mini prototype for further analysis. The
same process that was applied to the FeCrAl alloys was then
applied to the 304 steel, with the same conclusion: SS-2E is
the best geometry to mimic the SS-J type while also meeting Fig. 4. Engineering tensile curves in ‘‘plastic strain–engineering stress” coordinates
for alloy 6061-0 for specimens of different geometry. To demonstrate the PLC-effect
the criterion provided in Section 4.1.
peculiarities, the tensile curves were approximated and subtracted by the
polynomic curve.

4.4. Mechanical properties of the tested materials


complex, multi-scale background (Mazière and Forest, 2015). The
Table 3 demonstrates the average mechanical test results for specimen geometry change and the strong decrease in the number
the SS-J type geometry compared to the SS-2E for the tested mate- of grains involved in the process is expected to affect, at a mini-
rials. As shown in the table, for many materials the miniature mum, the kinetics of the serration flow and amplitude of the force
geometry provided close matching of the data between the two drops. However, the data shows not only good matching in the
types of specimen geometries. The typical difference in the yield mechanical properties (Table 3) but also excellent agreement in
and ultimate stress values did not exceed 5–8%, which is within the tensile behavior for SS-J and SS-Mini geometries (Fig. 4). The
the typical margin of error of a single measurement using either magnitude and type of the serrations were nearly identical.
geometry. Ductility values had slightly higher scattering, but, nev- Most alloys analyzed above demonstrate some ductility and
ertheless, the matching was deemed reasonable for most alloy fracture after demonstrating significant area reduction. In contrast,
development efforts and comparative studies. tungsten usually has very limited, if any, ductility and shows a brit-
Aluminum 6061 alloy is of special interest in research reactor tle fracture at room temperature. Small inaccuracies in specimen
core structures to form the skeleton, cladding, or the fuel or absor- and grips geometry can lead to minor deviations in the loading
ber matrix (Kim et al., 2008). Due to its good fabricability, excellent axis, which may become particularly problematic with specimen
corrosion performance (with appropriate pH control), high miniaturization. As outlined in Table 3, the SS-Mini 2E-0.4 geome-
strength in the low-temperature water coolant of these reactors try provided smaller fracture stress values (up to 25%) for tung-
(60–80 °C), and low neutron absorption rate, this material is rou- sten, compared to the SS-J type geometry. As believed, this issue
tinely used in the fuel and control components of these cores. Dur- may appear for brittle materials in general, not for tungsten only.
ing straining at room temperature with a conventional strain rate If so, special attention should be paid on the small deviations in
(10 3 s 1), this alloy after annealing (6061-0) demonstrated a the specimen geometry, surface quality, and other factors influenc-
specific phenomenon known as the Portevin-Le-Chatelier effect ing the test results.
(PLC). The PLC effect causes deformation localization and consecu-
tive formation of the localized shear bands. Numerous force drops
appeared in the tensile curves, as seen in Fig. 4. 4.5. Geometry limitations due to non-traditional scale factors
The evolution of the shear bands and amplitude of the force
drops depend on the collective behavior of grains in the polycrys- As discussed earlier, the grain size is a natural scale and allows
talline aggregate. Detailed analysis of this phenomena reveals a for estimating specimen size in ‘‘dimension-less” or ‘‘natural”
304 M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308

units. Such a ‘‘natural scale” may be found in some other cases, for UAM material, the grain size is not the dominating factor, and
instance, in the materials produced by ultrasonic additive manu- the observed properties of the 3D-printed material might be more
facturing (UAM) (Dehoff and Babu, 2010). The UAM technique sensitive to the specimen geometry and thickness, compared to a
exploits ultrasonic vibrations to effect a solid-state bonding conventional bulk alloy.
between metal tapes of 100–150 mm thickness, yielding a
multi-layer structure with numerous welded interfaces. The 4.6. Some methodological aspects of DIC tensile testing
UAM-produced materials have strong property inhomogeneity
depending on the loading direction (relative to the welding inter- If a set of images has been taken during the test and a reference
faces) (Schick et al., 2010). The direction perpendicular to the (non-deformed) image is available, strain field evolution may be
welding interfaces (often named ‘‘Z-direction”) is usually the analyzed using DIC algorithms (Sutton et al., 2009). The most com-
weakest one; Z-direction loading often leads to a quasi-brittle mon goals of DIC use are the strain field measurement (Fonseca
crack propagation and fracture with zero uniform and total ductil- et al., 2005; Sutton et al., 2009; Ambriz et al., 2013), analysis of
ity values. The UAM-produced tensile bars studied here had 7 property gradients in weldments (Field et al., 2014b), investigation
and 3 metal tape layers across the thickness of the SS-J and SS- of constitutive behavior and true stress–true strain curves (Džugan
2E specimens, respectively. et al., 2014; Kamaya and Kawakubo, 2014), dealing with phenom-
To estimate the role of the welding interface on a scale factor, ena such as Luders bands (Zhu et al., 2015), and investigating
mechanical tests are conducted with tensile specimens produced deformation localization and necking (Suzuki et al., 2010). Among
from UAM blocks (Fabrisonic, Columbus, OH) made of aluminum these, true curves and constitutive behavior parameters are espe-
6061-H18 alloy. The details of the UAM fabrication, bulk part pro- cially important for designing nuclear reactor components
duction, structure, and role of direction (X, Y, Z) on mechanical (Christopher et al., 2015).
behavior are given elsewhere (Sridharan et al., 2016) and will not In the mechanical test practice, plastic deformation in the spec-
be discussed here. Table 4 shows the mechanical properties of imen heads is often ignored; however, it may be of concern if there
the SS-J type geometry specimens compared to the SS-Mini 2E-0.4. is a need to perform focused ion beam (FIB) lift-out or produce
As shown in Table 3, the same bulk commercial aluminum alloy transmission electron microscopy (TEM) disks using the head
is almost insensitive to the scale factor; the results for the SS-J type material, which is a common practice for producing electron
and SS-Mini 2E-0.4 geometries are within reason. In contrast, the microscopy samples from irradiated material. Fig. 5 shows plastic
UAM-produced specimens demonstrate strong sensitivity to the strain distribution obtained by DIC analysis in 304L steel speci-
specimen geometry. In the case of SS-Mini 2E-0.4, the yield stress mens deformed at the overall strain level of 0.4. As shown in
is much smaller, and total elongation is much higher for X and Y the data, the strain level reaches about half of the overall strain
directions compared to the SS-J type. SS-Mini 2E-0.4 Z-direction level (0.2) observed at the gauge-head transition location (point
specimens have higher yield (fracture) stress than the SS-J type. A), and this value (eA) is close for all specimen geometries. At the
Most likely, this is caused by the probability of finding a welding distance of 0.8 mm from this location (point B), the strain level
flaw of a critical size capable of crack initiation and fracture. (eB) is still significant in the S-type specimen with reduced head
This limited study on the UAM objects highlights the need for size, whereas it drops to 0.03 in the E-type and SS-J specimen.
close inspection on the different scale factors contributing to the Note that even the SS-J specimen, with relatively massive heads
mechanical response of a specimen of interest. In the case of the (in relation to the miniature specimens under investigation), is

Table 4
The scale factor role on UAM-produced aluminum 6061 alloy.

Direction Specimen geometry Yield stress, MPa Ultimate stress, MPa Uniform elongation,% Total elongation,%
X – along the tape length SS-J type 202.9 206.9 0.5 6.8
SS-Mini (SS-2E) 176.9 216.7 0.7 8.5
Y – direction of straining during welding SS-J type 211.3 212.0 0.3 6.0
SS-Mini (SS-2E) 109.1 226.2 1.8 9.6
Z – perpendicular to the welding interfaces SS-J type 61.7 – – –
SS-Mini (SS-2E) 80.0 – – –

Fig. 5. Strain level (von Mises strain, Hencky’s strain tensor) in the head of the deformed specimens at a strain level of 0.4. Note the reduction in the local strain level in the
specimen head (from point A to point C for SS-J specimen geometry).
M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308 305

speckles of 3–7 pixels provide the best sensitivity and smallest


error (Sutton et al., 2009).

5. Test case for materials test reactor

It is unlikely that the effects of neutron irradiation experiments


could be fully simulated in the near future with tools like multiple
simultaneous ion-bombardment tests and advanced modeling and
simulation, even with substantial development activities ongoing
in both of these areas. If an irradiation project considers a new
specimen geometry, it may be advantageous to maintain compat-
ibility with existing infrastructure. To perform specimen irradia-
tion at HFIR, which is an 85 MWt beryllium-reflected,
pressurized, light-water-cooled and moderated flux-trap-type
Fig. 6. Typical true stress–true strain curves for specimens of annealed austenitic
reactor (Xoubi and Primm III, 2005), small cylindrical ‘‘rabbit” cap-
steel. Some increase in strain hardening rate is caused by martensitic
transformation. sules were developed over two decades ago to take advantage of
the very high neutron flux in this reactor (>1015 n/cm2 s in both
the thermal and fast regions of the spectrum). Each of these cap-
not large enough to completely eliminate plastic deformation in sules accommodated up to 24 SS-J type specimens. The general
the heads. As distance from the gauge increases, local strain internal capsule design remains the same from project to project,
decreases quickly (see point C for SS-J specimen in Fig. 5). For with only the gas gap value being adjusted to meet project-
the widest head portion, E- and SS-J geometries can provide virtu- specific requirements for the irradiation temperature.
ally non-deformed material. Note also that plastic strain in the In the present work, the rabbit capsule design is tailored to con-
heads should reduce quickly with strength level increase. tain various tensile specimen styles, as shown in Fig. 7. This versa-
Fig. 6 shows the true stress–true strain curves calculated using tile, modular configuration was developed to provide flexibility in
DIC for the annealed 304L steel. The measurements are conducted specimen loading and accommodates a useful range of irradiation
using a ‘‘virtual digital extensometer” for the middle portion of the temperatures in support of nuclear materials research. Each unique
specimen gauge. Here, annealed austenitic 304L steel is selected specimen configuration was designed to share the thermal equiva-
due to its relatively low yield stress and pronounced strain harden- lency, allowing the SS-J2 and SS-2E specimen modules to be inter-
ing. The difference between yield stress and ultimate stress reaches changeable with no impact on the design performance. This
3  as shown in Table 3, leading to increased stress in the heads equivalency was achieved by making the tensile specimens and
during straining. Nevertheless, as shown in the results from Fig. 6, their supporting parts (chevrons, holders, liners, etc.) form a uni-
the curves are practically identical for all specimen geometries. form ‘‘coupon” module with common mass, geometry, neutronic,
Small specimen geometries, even the ones with the smallest heads, and thermal properties. The internal components for the assembly
provide close matching. This result was expected (Džugan et al., configuration are shown in Fig. 8.
2014); DIC eliminates the inaccuracy caused by the plastic strain A key component of any irradiation program is to obtain the
in the heads. desired irradiation dose (in dpa and/or neutron fluence) and irradi-
In the context of the present work, annealed 304L steel with a ation temperature. To address the newly developed SS-2E geome-
strong strain hardening rate represents the worst case scenario. try impact on design temperatures, an ANSYS finite element
Most likely, as mentioned above, plastic deformation in the heads analysis (FEA) model is used. Given the modular nature of the
and its impact on the overall specimen behavior will be less pro- design, only a single tensile configuration layer located at the cap-
nounced in irradiated specimens and materials with smaller hard- sule centerline is analyzed to determine thermal performance.
ening rates such as the FeCrAl or 718-alloys. Scoping analyses are performed to show that all generic configura-
One methodological aspect worth noting here is that in the case tions of SS-J2/SS-2E tensile specimens perform in the same fashion,
of miniature specimens, creating DIC patterns may become an given the virtually identical heat loading and contact gaps seen in
issue. Traditional spray paint cans usually do not work well and the loading permutations. Table 5 outlines this thermal equiva-
only provide a moderate speckle pattern density (Berfield et al., lency for an irradiation running at a nominal temperature of
2007). Much denser patterns with smaller speckles can be created 550 °C. There is a larger span of temperatures for the SS-J specimen
with commercially available airbrushes, as noted by (Berfield et al., caused by the large temperature gradients observed in the grip sec-
2007). With airbrushes, the air pressure (and therefore the speckle tion of the specimens. However, the tensile gauge lengths for the
pattern density) can be adjusted to suit the video system resolu- specimens are more analogous, as seen in Fig. 9. Hence, the SS-
tion to get the optical pattern quality and data volume. Most often, 2E specimen geometry and the pre-existing capsule design may

Fig. 7. The HFIR rabbit capsule design for SS-J and SS-Mini tensile specimens. SiC: irradiation temperature sensor for SiC-thermometry.
306 M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308

Fig. 8. Specimens and modules prior to capsule assembly (left) and rabbit capsule internals (right). Also see Fig. 7.

Table 5
Design performance of the modular tensile design demonstrating thermal equivalence of the 550 °C temperature case.

Type Component Mass(g) Location Temperature (°C)


Avg. Min. Max
SS-2E module 6 SS-2E specimens 0.186 Inner 550 530 560
Middle 540 536 547
Outer 525 515 531
SS-J module 3 SS-J (T = 0.5 mm) 0.526 Inner 547 506 580
Middle 536 502 570
Outer 515 498 548

data—exists for the materials studied in Table 1. Given this, the


reduced mass of the miniaturized specimen geometry (nearly
7  less) should result in a significantly reduced radiological threat
on a per specimen basis. Clearly, follow-on work will address the
performance of the SS-2E geometry in the irradiated state, includ-
ing tensile performance, thermal performance, and radiological
threat performance.

6. Conclusions

The present work provides a complete engineering solution and


experimental database on a small specimen geometry suitable for
irradiation in materials test reactors and post-irradiation out-of-
hot cell testing. A sub-sized specimen geometry, ‘‘SS-Mini,” was
designed; the new geometry leverages existing geometries, and
configurations for irradiation capsules lead to reduced cost for
deployment. The performance of the proposed geometry was eval-
uated using 304L stainless steel and several commercial alloys
Fig. 9. Analyzed temperature comparison between SS-J2 and SS-2E specimens (°C). with several different strength and ductility levels. Mechanical ten-
sile tests were carried out to compare the engineering mechanical
properties (yield stress, ultimate tensile stress, and uniform and
total elongation values) and plastic behavior of different miniature
be used in all flux trap positions available in the HFIR. Furthermore,
specimens. It is shown that the proposed specimen geometry pro-
the modular subassemblies provide inherent specimen grouping
vides acceptable mechanical property results and meets the design
and simplify the post-irradiation disassembly. This feature pro-
parameters for use in a materials test reactor.
vides valuable cost savings, given the relatively high cost of hot cell
facility usage.
As stated previously, the design of the miniature test specimen Acknowledgements
should have reduced mass to limit the activated volume during
neutron irradiation. Because the SS-2E specimen geometry is This research was primarily funded by the U.S. Department of
newly developed, no irradiation data—and hence no activation Energy, Office of Nuclear Energy, for the Nuclear Energy Enabling
M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308 307

Technologies (NEET) program for the Reactor Materials effort. Lab- Jaya, N.B., Alam, M.Z., 2013. Small-scale mechanical testing of materials. Curr. Sci.
105, 1073–1099.
oratory Directed R&D funds at ORNL were used for testing of UAM
Jiang, J., Britton, T., Wilkinson, A., 2013. Evolution of dislocation density
manufactured aluminum. The authors would like to thank Dr. L.M. distributions in copper during tensile deformation. Acta Mater. 61, 7227–7239.
Garrison (ORNL) for providing tungsten tensile specimens for the Jung, P., Hishinuma, A., Lucas, G., Ullmaier, H., 1996. Recommendation of
present work and D.P. Stevens and S. Crawford (ORNL) for help miniaturized techniques for mechanical testing of fusion materials in an
intense neutron source. J. Nucl. Mater. 232, 186–205.
with manuscript preparation. Kamaya, M., Kawakubo, M., 2014. True stress–strain curves of cold worked stainless
steel over a large range of strains. J. Nucl. Mater. 451, 264–275.
Kaneda, J., Koshiishi, M., Morra, M.M., Rebak, R.B., et al., 2011. Microstructural
References characterization of surface modified alloy 82 welds regarding susceptibility to
environmentally assisted cracking. In: CORROSION2011. NACE International.
Algarni, M., Bai, Y., Choi, Y., 2015. A study of Inconel 718 dependency on stress Keller, C., Hug, E., Retoux, R., Feaugas, X., 2010. TEM study of dislocation patterns in
triaxiality and Lode angle in plastic deformation and ductile fracture. Eng. Fract. near-surface and core regions of deformed nickel polycrystals with few grains
Mech. 147, 140–157. across the cross section. Mech. Mater. 42, 44–54.
Alsabbagh, A., Valiev, R.Z., Murty, K., 2013. Influence of grain size on radiation Kim, J.W., Byun, T.S., 2010. Analysis of tensile deformation and failure in austenitic
effects in a low carbon steel. J. Nucl. Mater. 443, 302–310. stainless steels: Part I-Temperature dependence. J. Nucl. Mater. 396, 1–9.
Ambriz, R., Froustey, C., Mesmacque, G., 2013. Determination of the tensile behavior Kim, Y.S., Hofman, G., Robinson, A., Snelgrove, J., Hanan, N., 2008. Oxidation of
at middle strain rate of AA6061-T6 aluminum alloy welds. Int. J. Impact Eng. 60, aluminum alloy cladding for research and test reactor fuel. J. Nucl. Mater. 378,
107–119. 220–228.
Ayres, R.A., 1985. Thermal gradients, strain rate, and ductility in sheet steel tensile Klecka, J., Di Gabriele, F., Hojna, A., 2015. Mechanical properties of the steel T91 in
specimens. Metall. Trans. A 16, 37–43. contact with lead. Nucl. Eng. Des. 283, 131–138.
Bažant, Z.P., 1999. Size effect on structural strength: a review. Arch. Appl. Mech. 69, Klueh, R., 1985. Miniature tensile test specimens for fusion reactor irradiation
703–725. studies. Nucl. Eng. Des. Fusion 2, 407–416.
Berfield, T., Patel, J., Shimmin, R., Braun, P., Lambros, J., Sottos, N., 2007. Micro-and Kohno, Y., Kohyama, A., Hamilton, M.L., Hirose, T., Katoh, Y., Garner, F.A., 2000.
nanoscale deformation measurement of surface and internal planes via digital Specimen size effects on the tensile properties of JPCA and JFMS. J. Nucl. Mater.
image correlation. Exp. Mech. 47, 51–62. 283, 1014–1017.
Bleys, P., Kruth, J.-P., Lauwers, B., Schacht, B., Balasubramanian, V., Froyen, L., Van Kohyama, A., Hamada, K., Matsui, H., 1991. Specimen size effects on tensile
Humbeeck, J., 2006. Surface and sub-surface quality of steel after EDM. Adv. properties of neutron-irradiated steels. J. Nucl. Mater. 179, 417–420.
Eng. Mater. 8, 15–25. Mazière, M., Forest, S., 2015. Strain gradient plasticity modeling and finite element
Byun, T.S., Kim, J.H., Chi, S.H., Hong, J.H., 1998. Effect of Specimen Thickness on the simulation of Lüders band formation and propagation. Continuum Mech.
Tensile Deformation Properties of SA508 Cl. 3 Reactor Pressure Vessel Steel. In: Thermodyn. 27, 83–104.
Small Specimen Test Techniques. ASTM International. Panayotou, N., Atkin, S., Puigh, R., Chin, B., 1986. Design and use of nonstandard
Chen, M., Terada, D., Shibata, A., Tsuji, N., 2013. Identical area observations of tensile specimens for irradiated materials testing. In: The Use Small-Scale
deformation-induced martensitic transformation in SUS304 austenitic stainless Specimens Testing Irradiated Material. ASTM International.
steel. Mater. Trans. 54, 308–313. Parish, C.M., Kumar, N.K., Snead, L.L., Edmondson, P.D., Field, K.G., Silva, C., Williams,
Christopher, J., Choudhary, B., Kumar, R.V., Karthik, V., 2015. Constitutive A.M., Linton, K., Leonard, K.J., 2015. LAMDA: irradiated-materials microscopy at
description of flow behaviour of post-irradiated type 316 austenitic stainless Oak Ridge National Laboratory. Microsc. Microanal. 21, 1003–1004.
steel at low dpa. Nucl. Eng. Des. 291, 163–167. Pierron, O., Koss, D., Motta, A., 2003. Tensile specimen geometry and the
Connolley, T., Mchugh, P., Bruzzi, M., 2005. A review of deformation and fatigue of constitutive behavior of Zircaloy-4. J. Nucl. Mater. 312, 257–261.
metals at small size scales. Fatigue Fract. Eng. Mater. Struct. 28, 1119–1152. Rickerby, D., Fenici, P., 1986. Ductility of thin sheet tensile specimens for irradiation
Dai, Y., Bauer, G., 2001. Status of the first SINQ irradiation experiment, STIP-I. J. Nucl. experiments. Nucl. Eng. Des. Fusion 3, 423–424.
Mater. 296, 43–53. Rund, M., Procházka, R., Konopik, P., Džugan, J., Folgar, H., 2015. Investigation of
Dehm, G., Motz, C., Scheu, C., Clemens, H., Mayrhofer, P.H., Mitterer, C., 2006. sample-size influence on tensile test results at different strain rates. Proc. Eng.
Mechanical size-effects in miniaturized and bulk materials. Adv. Eng. Mater. 8, 114, 410–415.
1033–1045. Schick, D., Hahnlen, R., Dehoff, R., Collins, P., Babu, S., Dapino, M., Lippold, J., 2010.
Dehoff, R., Babu, S., 2010. Characterization of interfacial microstructures in 3003 Microstructural characterization of bonding interfaces in aluminum 3003
aluminum alloy blocks fabricated by ultrasonic additive manufacturing. Acta blocks fabricated by ultrasonic additive manufacturing-methods were
Mater. 58, 4305–4315. examined to link microstructure and linear weld density to the mechanical
Džugan, J., Procházka, R., Konopik, P., 2014. Micro-Tensile Test Technique properties of ultrasonic additive manufacturing. Weld. J. 89, 105S.
Development and Application to Mechanical Property Determination. In: Shin, C., Lim, S., Jin, H., Hosemann, P., Kwon, J., 2015. Specimen size effects on the
Small Specimen Test Techniques 6th Volume. ASTM International. weakening of a bulk metastable austenitic alloy. Mater. Sci. Eng., A 622, 67–75.
Fazio, C., Briceno, D.G., Rieth, M., Gessi, A., Henry, J., Malerba, L., 2011. Innovative Shirdel, M., Mirzadeh, H., Parsa, M., 2015. Estimation of the kinetics of martensitic
materials for Gen IV systems and transmutation facilities: the cross-cutting transformation in austenitic stainless steels by conventional and novel
research project GETMAT. Nucl. Eng. Des. 241, 3514–3520. approaches. Mater. Sci. Eng., A 624, 256–260.
Field, K.G., Gussev, M.N., Busby, J.T., 2014a. Microstructural characterization of Signor, L., Villechaise, P., Ghidossi, T., Lacoste, E., Gueguen, M., Courtin, S., 2016.
deformation localization at small strains in a neutron-irradiated 304 stainless Influence of local crystallographic configuration on microcrack initiation in
steel. J. Nucl. Mater. 452, 500–508. fatigued 316LN stainless steel: Experiments and crystal plasticity finite
Field, K.G., Gussev, M.N., Yamamoto, Y., Snead, L.L., 2014b. Deformation behavior of elements simulations. Mater. Sci. Eng., A 649, 239–249.
laser welds in high temperature oxidation resistant Fe–Cr–Al alloys for fuel Sridharan, N., Gussev, M., Seibert, R., Parish, C., Norfolk, M., Terrani, K., Babu, S.S.,
cladding applications. J. Nucl. Mater. 454, 352–358. 2016. Rationalization of anisotropic mechanical properties of Al-6061
Fonseca, J..Quinta., P., Withers, P.,, da, Mummery., 2005. Full-field strain mapping by fabricated using ultrasonic additive manufacturing. Acta Mater. 117, 228–
optical correlation of micrographs acquired during deformation. J. Microsc. 218, 237.
9–21. Sutton, M.A., Orteu, J.J., Schreier, H., 2009. Image correlation for shape, motion and
Greer, J.R., De Hosson, J.T.M., 2011. Plasticity in small-sized metallic systems: deformation measurements: basic concepts, theory and applications. Springer
Intrinsic versus extrinsic size effect. Prog. Mater Sci. 56, 654–724. Science & Business Media.
Gussev, M., Busby, J., Field, K., Sokolov, M., Gray, S., 2014a. Role of scale factor during Suzuki, K., Jitsukawa, S., Okubo, N., Takada, F., 2010. Intensely irradiated steel
tensile testing of small specimens. In: Small Specimen Test Techniques 6th components: plastic and fracture properties, and a new concept of structural
Volume. ASTM International. design criteria for assuring the structural integrity. Nucl. Eng. Des. 240, 1290–
Gussev, M.N., Field, K.G., Busby, J.T., 2014b. Strain-induced phase transformation at 1305.
the surface of an AISI-304 stainless steel irradiated to 4.4 dpa and deformed to Viehrig, H.-W., Altstadt, E., Houska, M., 2015. Radiation response of the overlay
0.8% strain. J. Nucl. Mater. 446, 187–192. cladding from the decommissioned WWER-440 Greifswald unit 4 reactor
Hindley, M.P., Blaine, D.C., Groenwold, A.A., Becker, T.H., 2015. Failure prediction of pressure vessel. Nucl. Eng. Des. 286, 227–236.
full-size reactor components from tensile specimen data on NBG-18 nuclear Wakai, E., Nogami, S., Kasada, R., Kimura, A., Kurishita, H., Saito, M., Ito, Y., Takada,
graphite. Nucl. Eng. Des. 284, 1–9. F., Nakamura, K., Molla, J., et al., 2011. Small specimen test technology and
Hoffmann, H., Hong, S., 2006. Tensile test of very thin sheet metal and methodology of IFMIF/EVEDA and the further subjects. J. Nucl. Mater. 417,
determination of flow stress considering the scaling effect. CIRP Ann. Manuf. 1325–1330.
Technol. 55, 263–266. Wang, C., Wang, C., Guo, B., Shan, D., Huang, G., 2013. Size effect on flow stress in
Howard, C., Frazer, D., Lupinacci, A., Parker, S., Valiev, R., Shin, C., Choi, B.W., uniaxial compression of pure nickel cylinders with a few grains across
Hosemann, P., 2016. Investigation of specimen size effects by in-situ thickness. Mater. Lett. 106, 294–296.
microcompression of equal channel angular pressed copper. Mater. Sci. Eng., Was, G., Busby, J., Allen, T., Kenik, E., Jensson, A., Bruemmer, S., Gan, J., Edwards, A.,
A 649, 104–113. Scott, P., Andreson, P., 2002. Emulation of neutron irradiation effects with
Hurst, R.C., Matocha, K., 2015. A Renaissance in the use of the small punch testing protons: validation of principle. J. Nucl. Mater. 300, 198–216.
technique. In: ASME2015 Pressure Vessels Piping Conference. American Society Xoubi, N., Primm III, R., 2005. Modeling of the High Flux Isotope Reactor Cycle 400.
of Mechanical Engineers, pp. V01BT01A048-V01BT01A048. ORNL/TM-2004/251. Oak Ridge National Laboratory, Oak Ridge, Tennessee.
308 M.N. Gussev et al. / Nuclear Engineering and Design 320 (2017) 298–308

Yamamoto, Y., Pint, B., Terrani, K., Field, K., Yang, Y., Snead, L., 2015. Development Yvon, P., Le Flem, M., Cabet, C., Seran, J.L., 2015. Structural materials for next
and property evaluation of nuclear grade wrought FeCrAl fuel cladding for light generation nuclear systems: Challenges and the path forward. Nucl. Eng. Des.
water reactors. J. Nucl. Mater. 467, 703–716. 294, 161–169.
Yang, B., Liaw, P., Wang, G., Morrison, M., Liu, C., Buchanan, R., Yokoyama, Y., 2004. Zhang, Y., Topping, T.D., Lavernia, E.J., Nutt, S.R., 2014. Dynamic micro-strain
In-situ thermographic observation of mechanical damage in bulk-metallic analysis of ultrafine-grained aluminum magnesium alloy using digital image
glasses during fatigue and tensile experiments. Intermetallics 12, 1265–1274. correlation. Metall. Mater. Trans. A 45, 47–54.
Yang, B., Liu, C., Nieh, T., Morrison, M., Liaw, P., Buchanan, R., 2006. Localized Zhu, F., Bai, P., Zhang, J., Lei, D., He, X., 2015. Measurement of true stress–strain
heating and fracture criterion for bulk metallic glasses. J. Mater. Res. 21, 915– curves and evolution of plastic zone of low carbon steel under uniaxial tension
922. using digital image correlation. Opt. Lasers Eng. 65, 81–88.
Yuan, W., Zhang, Z., Su, Y., Qiao, L., Chu, W., 2012. Influence of specimen thickness
with rectangular cross-section on the tensile properties of structural steels.
Mater. Sci. Eng., A 532, 601–605.

Das könnte Ihnen auch gefallen