Sei sulla pagina 1di 9

Experimental Mechanics (2010) 50:37–45

DOI 10.1007/s11340-008-9210-9

Nanomechanical Testing of Gum Metal


E.A. Withey & J. Ye & A.M. Minor & S. Kuramoto &
D.C. Chrzan & J.W. Morris Jr.

Received: 28 May 2008 / Accepted: 1 December 2008 / Published online: 24 January 2009
# Society for Experimental Mechanics 2009

Abstract “Gum Metal” is a newly developed β-Ti alloy The pattern of deformation is consistent with prior work
that, in the cold-worked condition, has exceptional elastic showing deformation by the formation and growth of shear
elongation and high strength. The available evidence bands and faults in a matrix that is densely decorated with
suggests that Gum Metal does not yield until the applied defects.
stress approaches the ideal strength, and then deforms by
mechanisms that do not involve conventional crystal Keywords Gum metal . Titanium . Sub-micron pillars .
dislocations. To study its behavior, submicron-sized pillars TEM . Compression
of solution-treated and cold-worked Gum Metal were
compressed in situ in a quantitative compression stage in
a transmission electron microscope. Solution-treated speci- Introduction
mens and half of the cold-worked specimens exhibited
essentially monotonic hardening during compression, but The ideal strength of a material is reached when its atomic
with serrated load-deflection curves that included periodic bonds become so stretched that they can no longer support
partial relaxations of the stress. The other cold-worked the applied load. At that point of elastic instability the
specimens exhibited pronounced shear instability. These material must fracture or deform. The ideal strength in
samples deformed by a stick-slip motion along a well- shear can be computed from first principles and, for a body-
defined shear plane, with a serrated load-deflection curve centered-cubic (bcc) metal, is about 11% of the shear
demonstrating partial stress relaxation at each sliding event. modulus [1]. However, most metals yield at stresses
significantly lower than their ideal strength due to the
motion of dislocations. In order for a metal to reach its ideal
strength, the dislocations within it must be so efficiently
E.A. Withey (*) : J. Ye : A.M. Minor : D.C. Chrzan : pinned that they cannot move until the ideal strength is
J.W. Morris Jr.
exceeded.
Department of Materials Science and Engineering,
University of California, In 2003, researchers at Toyota Central R&D introduced
Berkeley, CA 94720, USA an intriguing new class of β-titanium alloys, coined Gum
e-mail: ewithey@berkeley.edu Metal, that after severe cold-working appear to deform at,
J. Ye : A.M. Minor
or near, their ideal strength [2]. These alloys were designed
National Center for Electron Microscopy, to have a near-zero shear modulus in the <111> direction
Lawrence Berkeley National Laboratory, [3], the weakest direction in the bcc structure [1, 4], and
Berkeley, CA 94720, USA have a generic composition Ti–24(Nb+Ta+V)–(Zr+Hf)–O
(at.%) [5]. Examples of Gum Metal include Ti–12Ta–9Nb–
S. Kuramoto
Toyota Central R&D Laboratories, Inc., 3V–6Zr–1.2O, Ti–23Nb–0.7Ta–2Zr–1.2O, Ti–20Nb–
Nagakute, Aichi 480-1192, Japan 3.5Ta–3.4Zr–1.2O (at.%). After extreme cold-working
38 Exp Mech (2010) 50:37–45

(∼90%), these alloys display a number of “super” properties Metal is on the order of 10 nm. Thus, the microstructure is,
including super elasticity, high strength, super cold- at least, generally consistent with the requirements for a
workability, and Invar (low thermal expansion coefficient material to reach its ideal strength. However, the nature of
over a temperature range) and Elinvar (constant elastic the obstacles that pin the dislocations remains unclear.
modulus over a limited temperature range) properties [2]. The domain structure inside the nanoindentation pits
Gum Metal is currently used in eyeglass frames, as studied in [7] mimicked the microstructure of the material
precision screws and fishing line, and in tennis rackets in the CW condition. Examination of this structure in both
and golf clubs. Given its low elastic modulus and high bulk and deformed CW and deformed ST Gum Metal
strength, future biomedical, automotive, and aeronautical samples has shown that there are slight rotations from one
applications are expected. domain to the next suggesting that there is a nearly
In the solution-treated (ST) condition, Gum Metal has a continuous rotation across multiple fuzzy domains. The
microstructure and a set of properties that resemble other structure of the boundaries between domains is not yet
bcc alloys: large equiaxed grains, ∼50–100 μm in diameter, clear.
and a stress–strain curve with linear–elastic behavior In the work reported here the mechanism of plastic
followed by yielding. In contrast, cold-worked (CW) Gum deformation of Gum Metal in the ST and CW conditions
Metal has a marble-like optical microstructure, which, on a was investigated by compressing sub-micron-sized pillars
smaller scale, appears as fuzzy domains ∼200 nm in of Gum Metal in an indentation stage mounted in a
diameter with ill-defined boundaries each rotated slightly transmission electron microscope (TEM). This technique
with respect to its neighbors. The tensile stress-strain curve was chosen because it allows the deformation process to be
includes a large elastic deformation that becomes non-linear followed in situ in samples whose small size ensures that
before yielding. The plastic elongation is appreciable, but internal stresses are reduced and substantial heterogeneities
includes very little hardening. are avoided.
According to Gutkin et al., the formation of the marble-
like structure develops through the formation of and shear
along “giant” faults, which are defined planar regions Experimental Procedure
∼1 nm thick within which localized plastic strain reaches
thousands of percent [6]. These defects were first seen by Gum Metal with a composition of Ti–35.9Nb–2Ta–2.7Zr–
Saito et al. while pulling ST Gum Metal in uniaxial tension 0.3O (wt.%) was obtained from Toyota Central R&D
to 10% strain [2]. The direction of the faults was ∼45° from Laboratories, Nagoya, Japan in the form of two 4 mm
the tensile axis, but did not parallel typical slip planes or diameter rods. Each rod was processed from elemental
slip directions for the bcc structure. Subsequent investiga- powders cold isostatically pressed into cylindrical form
tion of giant faults and their role in developing the before sintering at 1,300°C and hot forging at 1,150°C in
microstructure of CW Gum Metal showed that as the air, and solution-treating at 900°C in Ar. Rods were then
amount of work increased, the faults became visible across cold-rotary swaged by 90%. One rod was subsequently
the large grains of the ST Gum Metal [5, 6]. Giant faults solution-treated a second time at 900°C for 30 min, thus
were not observed to penetrate the high-angle grain producing one CW or cold-rotary swaged rod and one ST or
boundaries, but did produce steps in the grain boundaries solution-treated rod [9].
they intersected. Near the giant faults, the orientations of Sections of each rod were cut parallel to the swaging
the grains were rotated from their original orientations. As direction (rod axis) with thicknesses of ∼500 μm. These
the amount of work increased further, the grains became sections were mechanically thinned to ∼20 μm and attached
more elongated and serrated eventually leading to the with super glue to a copper substrate. They were then
marble-like structure seen in 90% CW Gum Metal. machined in a FEI 235 dual focused ion beam (FIB) using
Prior work in this laboratory (Withey et al. [7]) on the an annular pattern and low current (∼10 nA) to produce
nanoindentation of one composition of Gum Metal in cylindrical pillars 100–200 nm in diameter. Each pillar had
the ST condition revealed thick arrays of dislocations in a sidewall taper of 4–5° defined by the angle of the sidewall
the peripheries of the indentation pits, while the fuzzy with respect to the axis of the pillar, and length up to 1 μm.
domains typical of cold-worked material filled the severely The copper substrate with the attached Gum Metal
deformed volume along the sides and bottoms of the pits. pillars was placed in a Hysitron PicoIndenter® stage in
The dislocations were dipoles connected by tightly serrated a JEOL 3010 TEM located in the National Center for
end segments indicating the presence of nanoscaled Electron Microscopy (NCEM) at the Lawrence Berkeley
obstacles spaced ∼25 nm apart. Parallel theoretical work National Laboratory. The PicoIndenter® is equipped with a
(Li et al. [8]) found that the obstacle spacing required to pin boron-doped diamond flat punch tip with an area much
dislocations at stresses above the ideal strength of Gum larger than the contact areas of the pillars. The indentation
Exp Mech (2010) 50:37–45 39

Results

During the compression of both ST and CW pillars, stress


and deformation was localized at the tops of the pillars
apparently due to their sidewall tapers. All pillars had a
speckled appearance before compression, which was
ascribed to damage on the outside layer of the pillars due
to FIB processing. The density of these speckles differed
with local orientation of the pillars.
The basic results of the compression experiments are
documented in Figs. 2, 3 and 4, which illustrate the two
distinct behavioral patterns that were observed. Each figure
compares load–displacement curves for two metallurgically
similar samples that exhibited similar behavior. In each
case, the deformation of the two samples is compared at
roughly similar stages of overall deformation by extracting
frames from the test films. The three frames were taken at
the points indicated on the load deflection curves. The
deformation patterns are compared directly in Fig. 5.
Figure 5(a) shows sample load deflection curves corre-
sponding to the superimposed traces of the sample profiles

Fig. 1 (a) Indentation stage and (b) TEM image of the flat punch and
an ST pillar after a compression test showing the large size of the
punch compared to the pillar

stage and flat punch are shown in Fig. 1(a) and (b),
respectively. Load and displacement resolutions were,
nominally, 0.1 μN and 0.5 nm, respectively. More detailed
information about the PicoIndenter® can be found in [10].
Before each compression experiment, diffraction patterns
were taken from several locations on the pillars to charac-
terize their initial microstructure and orientation to be
compared with their microstructure and orientation after
compression. Each compression test was done under
displacement control at a displacement rate of 20 nm/s in
either bright-field or dark-field imaging modes. The com-
pressions were recorded in real time at 30 frames per second
by a camera within the microscope. Load–displacement data
was measured continuously throughout each test by a
Fig. 2 (a) Load–displacement data for two ST pillars with the black
feedback loop in the indentation stage, and matched frame- data corresponding to the frames extracted in (b–d) and the grey data
by-frame with test film. to those in (e–g)
40 Exp Mech (2010) 50:37–45

region of deformation which is defined by a scallop on


the edge of the deformed pillar. The diffraction patterns of
the deformed regions of the ST pillars [Fig. 6(b)] are
streaked to form a broken ring pattern about the <110> zone
axis, in contrast to the sharp diffraction spots in the <110>
zone axis orientation of the pillar before deformation
[Fig. 6(a)] indicating the formation of several new domains.
The deformed regions are, slightly rotated from their original
orientation, as was previously observed in studies of ex-situ
nanoindentation [7].
Figure 2(e–g) shows the comparable behavior for the
other ST pillar. In this case multiple streaks also appear, but
are roughly perpendicular to the pillar axis. Given the
similarity of the two load-deflection curves, we believe that
the slip lines are at 45° to the axis in this case as well, but
are oriented into the plane of the figure.
The serrated load-deflection curves measured on these
specimens are not unusual. However, the serrations
observed here are significantly less pronounced than those
seen in prior pillar compression tests of more conventional
crystalline materials [10]. The pronounced load drops
observed in the prior tests could usually be associated with

Fig. 3 (a) Load–displacement data for two non-sliding CW pillars


with the black data corresponding to the frames extracted in (b–d) and
the grey data to those in (e–g)

shown in Fig. 5(b–d) taken at three similar points along the


load deflection curves.
Figures 2 and 3 show sets of ST and CW specimens
that have roughly similar behavior. The load–displacement
curves have a serrated stair-step shape that shows a general
hardening trend with periodic load drops.

Solution-Treated (ST) Specimens

The load-deflection curves of the two ST pillars virtually


superimpose for the first 80 nm of displacement [Fig. 2(a)].
The general hardening trend is punctuated by a few
significant load drops interspersed with regions of slight
softening where the load appears to decrease linearly with
increasing displacement before increasing to another local
maximum. The softening portion of the curve persists over
about 15–20 nm of displacement, and terminates in a well-
defined minimum. Figure 2(b–d) and (e–g) shows the
corresponding frames of the video for each of the two
pillars. In Fig. 2(b–d), as the flat punch compresses the pillar,
four dark streaks form at approximately 45° to the pillar Fig. 4 (a) Load–displacement data for two non-sliding CW pillars
axis. Closer investigation of still images taken after with the black data corresponding to the frames extracted in (b–d) and
compression [Fig. 6(b)], show that each streak borders a the grey data to those in (e–g)
Exp Mech (2010) 50:37–45 41

Fig. 5 (a) Load displacement curves for ST, non-sliding CW, and sliding CW pillars indicating the points from which the traces of the respective
pillars are shown in (b), (c), and (d)

bursts of dislocation activity, producing a spontaneous The load-deflection curves of two CW samples that
stress relaxation that essentially relieved the imposed load. exhibited monotonic hardening are shown superimposed in
Since no dislocation motion was directly observed during Fig. 3 along with captured frames that show the change in
the compression of the ST pillars, it is difficult to say shape during deformation. A sequential trace of the profile
whether the load drops seen here were due to dislocation of the specimen associated with the black curve in Fig. 3 is
mobility or to some other mechanism. However, the sample shown in Fig. 5(c). Before and after pictures of the
does not relax in obvious bursts; the relaxation is gradual specimen that produced the load-deflection curve that is
and stops while the sample remains under significant load. colored gray in Fig. 3 are shown in Fig. 7.
The deformation mechanism appears to involve the gradual As in the case of the solution-treated material, the two
growth of planar “shear bands” against a significant resis- load-deflection curves almost superimpose. As in the ST
tance that leads to an overall monotonic hardening. case, the curves are serrated. But the serrations are much
less pronounced; the load drops are smaller. Deformation is
Cold-Worked (CW) Specimens: Monotonic Hardening narrowly confined to the sample tip. As the indentation
proceeds, the pillar slowly and progressively develops a
The cold-worked specimens exhibited two very different sheared “foot”. The slow development of the deformation is
patterns of behavior during pillar compression: monotonic evident from the superimposed sample profiles in Fig. 5(c);
hardening (Fig. 3) and “shear instability” (Fig. 4). the sheared tip is most apparent in the post-test micrograph

Fig. 6 TEM images taken


(a) before and (b) after com-
pression of the ST pillar shown
in Fig. 2(b–d). The top diffrac-
tion pattern in (b) corresponds
to the deformed volume of the
pillar and the bottom to the
undeformed volume. The dif-
fraction pattern in (a) gives the
orientation of the pillar
before deformation
42 Exp Mech (2010) 50:37–45

Fig. 7 TEM images taken (a) before and (b) after compression of the CW pillar shown in Fig. 3(e–g). The top diffraction pattern in (a)
corresponds to the first third of the pillar from the tip and the bottom pattern from the second third. In (b) the top pattern corresponds to the
deformed volume of the pillar and the bottom pattern to the volume right below the deformed region

in Fig. 7(b). The diffraction patterns included in Fig. 7 along with the relevant diffraction patterns. The second
show that this sample contained at least two crystallographic sample, in Fig. 4(e–g), appears to have deformed by a
domains. The upper domain is severely deformed, and its similar shear instability, but in this case the plane of shear is
diffraction pattern shows spots that are slightly streaked. angled into the plane of the figure and was difficult to
The lower domain appears unchanged. The profile of the image in the TEM.
sheared tip bears a general resemblance to the final The load-deflection curves for these specimens are
profile of the ST sample in Fig. 6(b), but the boundary serrated, like the curves for the ST specimens in Fig. 2,
between the sheared and unsheared material is much but differ in three important respects. First, the loads are
sharper in this case. significantly smaller (both ST pillars and the sliding CW
pillars had similar initial diameters), reflecting the ease of
Cold-Worked (CW) Specimens: Shear Sliding deformation along the well-defined shear band. Second,
there is little monotonic hardening. In fact, the sample in
The load-deflection curves of the two samples that Fig. 4(b–d), that develops the visible shear offset, softens in
exhibited shear sliding are plotted in Fig. 4, along with the later stages of its compression. This is presumably due
micrographs that were taken at critical points in the defor- to the decreasing area of contact on the plane of shear.
mation to illustrate its pattern. The pillar in Fig. 4(b–d) Third, the depths of the serrations are comparable to those
show an obvious shear instability; a section of the sample in the ST samples and significantly greater than those in
tip is displaced by sliding on a narrow shear band angled the CW samples in Fig. 3, but have a very different
∼45° to the pillar axis. The development of the offset is significance than in the ST pillars. In the ST samples the
shown in the superimposed traces in Fig. 5(d). The final successive minima in the load–displacement curves appear
configuration of the offset sample is shown in Fig. 8(b), to be associated with the formation of spatially discrete slip

Fig. 8 TEM images taken (a) before and (b) after compression of the CW pillar shown in Fig. 4(b–d). The top diffraction patterns in (a) and (b)
correspond to the top domain, the domain that slid, and the bottom patterns are taken from the second domain from the top of the pillar
Exp Mech (2010) 50:37–45 43

bands, while in the CW specimens they appear to be due to the compression of submicron-sized pillars of metallic glass
a frictional stick-slide motion on a narrow, well-defined led to the formation of a shear band along which the
shear plane. The effective friction that opposes this sliding majority of subsequent deformation seemed to occur
is, apparently, considerable since the sample supports a through the offset of the top of the pillar at an angle to
significant load even at the minima of its serrations. the base of the pillar. Like the sliding tip, the offset of the
The diffraction patterns in Fig. 8 show that the offset top of the metallic glass pillar moved jerkily in spurts
sample in Fig. 4(b–d) has multiple domains. The severe accompanied by large load drops.
deformation is confined to the domain at the sample tip.
The apparent planarity of the shear band and its orientation Hardening
near 45° from the tensile axis suggests that it is confined to
a single domain and does not cross the boundary. A second interesting feature of the compression behavior is
the frictional resistance to shear and the pronounced
hardening that follows it. In typical in situ TEM deforma-
Discussion tion experiments on crystalline materials, the deformation
occurs primarily through a series of bursts in which
Patterns of Deformation dislocations or showers of dislocations are generated and
flow relatively freely through the crystal. This is not the
The results show that the indentation of Gum Metal case in Gum metal. Even the most pronounced serrations
produces at least two distinct deformation patterns: the relax only part of the applied load, indicating a substantial
monotonic hardening pattern observed in the solution- resistance to shear, and the relaxation is followed by a
treated and two of the four cold-worked specimens, and pronounced and continuous hardening (the only apparent
the shear instability experienced by the other two cold- exception—the macroscopic softening of the CW pillar in
worked specimens. Fig. 4(b–d)—is almost certainly a geometric effect due
The reason for the different behaviors of the cold- to the decreasing area of contact in the shear plane). Since
worked specimens appears to be crystallographic. Figures 7 the diffraction patterns of the material in these submicron-
and 8 include micrographs of the specimen in Fig. 3(e–g), sized pillars are well-defined before deformation occurs,
which showed monotonic hardening, and the specimen in this behavior cannot be attributed to residual stresses or
Fig. 4(b–d), which exhibited the clearest shear instability, lattice distortions, even in the severely cold-worked
respectively. In both cases <110> and <112> directions are material.
in the plane of the figure, taken from the diffractions patterns Prior work [7] has shown that the matrix of ST Gum
near the sample tip. The {110} and {112} preferred slip metal is densely decorated with features that strongly resist
planes are, of course, perpendicular to these directions. dislocation glide. While the precise nature of these obstacles
Figure 8 shows that the well-defined slide plane of the pillar remains unclear, they appear to be responsible for the
is parallel (or nearly so) to the trace of the (110) close- suppression of dislocation activity in the cold-worked
packed plane, a preferred plane for the formation of a sharp material. Assuming that those obstacles are present and
shear band like the one that developed in the sample. active in these samples (though they are not resolved in the
Figure 7 does not include a similar plane for the mono- diffraction patterns) they provide an obvious source for the
tonically hardened CW pillar, though the post-test data, friction and hardening that occurs in these tests.
which includes the deformation-induced rotations of the
crystal axes, suggests that the severely deformed material at Deformation Mechanism
the tip has rotated so that the deformed region is bounded by
a {110} plane. There are three mechanisms that may be relevant to the
While this crystallographic explanation seems persua- deformation of Gum metal: dislocation plasticity, micro-
sive, it is not sufficient to explain all of the data. For shearing leading to shear bands, and phase transformations.
example, the ST sample shown in Fig. 2(b–d) also had a Dislocation plasticity plays some role in the deformation of
{110} plane oriented for easy shear, and the sample does the solution-treated material, but has not been observed in
appear to shear on this plane. But the shear appears on cold worked specimens. Microshears and shear bands are
multiple {110} planes, leading to a more diffuse deforma- reasonably well documented in both solution-treated and
tion that does not produce a sharp offset like that in CW cold-worked material [2, 6]. There is preliminary evidence
pillars that had distinct sliding. for phase transformation during deformation, both to
Moreover, work by Shan et al. [11] includes evidence nanoparticles of ω-phase and to a reversible orthorhombic
that crystallographic slip is not necessarily required to phase (Dye D, private correspondence, 2008). While these
produce the well-defined offsets seen here. In their study, likely play a role in the elastic deformation there is no
44 Exp Mech (2010) 50:37–45

clear evidence of their contribution to bulk plasticity On the other hand, as discussed above, the pattern of
(Khachaturyan AG, private correspondence, 2008). deformation suggests that deformation occurs by local shear
The presence and persistence of sample preparation rather than dislocation glide, which is consistent with
damage (FIB damage) prevented the direct observation of deformation at the limit of elastic stability.
the deformation mechanism during these tests. On the other
hand, the very persistence of that damage, which consists
largely of dislocation loops, indicates that dislocations are Conclusions
not mobile in these specimens. In pure monocrystalline
pillar specimens, the FIB damage is swept out of the Submicron-sized pillars of Gum Metal were compressed in
specimen in the earliest stags of compression [10]. Further, situ in a quantitative compression stage in a transmission
there is no clear evidence of ω-phase or orthorhombic electron microscope. The samples included solution-treated
phase in the post-test diffraction patterns, even after and heavily cold-worked Gum Metal; the latter condition
extensive deformation. It therefore appears that the defor- leads to and exceptionally high strength in bulk material
mation occurs primarily through the formation of micro- that may reach the ideal strength of the material. The load-
shears or shear bands, as previously observed for bulk deflection curves and deformation patterns of the samples
deformation [2, 5, 6]. The difficulty of developing planar show two rather different responses. The two solution-
shear in a microstructure that is densely decorated with treated specimens and two of the four cold-worked speci-
nanoparticles would then explain the significant friction mens exhibited essentially monotonic hardening during
that opposes deformation, particularly when no crystallo- compression, but with serrated load-deflection curves that
graphically preferred shear plane is well oriented for slip. included periodic, partial relaxations of the stress. The other
two cold-worked specimens exhibited pronounced shear
Strength instability, with the samples deforming by a stick–slip
motion along a well-defined shear plane, with a serrated
One of the most interesting features of Gum metal is its load-deflection curve with partial stress relaxation at each
very high strength, particularly in the cold worked state, sliding event. While the defects that produced deformation
which approaches the predicted ideal strength of the could not be observed directly, the pattern of deformation is
material. That strength is apparent in these pillar compres- consistent with prior work showing deformation by the
sion tests. The ST pillars reach near ideal strengths and the formation and growth of shear bands and faults in a matrix
CW pillars have strengths just over half of the measured that is densely decorated with defects. This pattern is
strengths of the ST pillars. consistent with deformation at the limit of strength.
If we estimate the yield strength of the pillars by However, in the CW pillars the estimated stresses at yield
dividing the load at the first indication of plastic deforma- are well below the predicted values at the limit of strength.
tion by the cross-sectional area of the pillar at its apex, the The low strength may be due to geometric or surface effects
highest yield strengths are those of the ST specimens that are not yet understood.
(2,900–3,600 MPa), followed by the monotonically hard-
ening CW samples (∼1,900 MPa), followed by the speci- Acknowledgements This research was supported by the National
Science Foundation under Grant DMR 0706554 and by Toyota Motor
mens that deform by planar sliding (1,600–1,900 MPa). Corporation under a grant to the University of California Berkeley.
This is opposite of bulk behavior where the CW material EW also acknowledges support from NSF through a graduate research
yields at a higher strength than the ST material. This fellowship. Research at the National Center for Electron Microscopy
suggests a stronger size effect in the solution-treated Gum was supported by the Scientific User Facilities Division of the Office
of Basic Energy Sciences, U.S. Department of Energy under Contract
Metal. It is strongly suspected that the lower yield strength # DE-AC02-05CH11231.
of the CW pillars is due to residual stresses retained in the
material by the domain structure. While the source of this
early yielding in the CW pillars is not clear, the pattern of References
deformation suggests that shear instabilities emanate from
free surfaces and propagate along well-defined planes. 1. Krenn CR, Roundy D, Morris JW, Cohen ML (2001) Ideal
Since the geometry of the sample and indenter are not strength of BCC metals. Mater Sci Eng A 319–321:111–119,
perfectly controlled, there is the possibility that large stress doi:10.1016/S0921-5093(01)00998-4
concentrations develop on first contact, generating local, 2. Saito T, Furuta T, Hwang JH, Kuramoto S, Nishino K, Suzuki N,
Chen R, Yamada A, Ito K, Seno Y, Nonaka T, Ikehata H,
planar instabilities that propagate, driven by the stress Nagasako N, Iwamoto C, Ikuhara Y, Sakuma T (2003) Multifunc-
concentration at the periphery. Resolution of this issue will tional alloys obtained via a dislocation-free plastic deformation
require further investigation. mechanism. Science 300:464–467, doi:10.1126/science.1081957
Exp Mech (2010) 50:37–45 45

3. Ikehata H, Nagasako N, Furuta T, Fukumoto A, Miwa K, Saito T 8. Li T, Withey E, Morris J, Nagasako N, Kuramoto S, Chrzan D
(2004) First-principles calculations for development of low elastic (2007) Ideal deformation in Gum Metal. Ti-2007 Science and
modulus Ti alloys. Phys Rev B Condens Matter Mater Phys Technology: Proceedings of the 11th World Conference on
70:174113, doi:10.1103/PhysRevB.70.174113 Titanium
4. Krenn CR, Roundy D, Cohen ML, Chrzan DC, Morris JW (2002) 9. Furuta T, Nishino K, Hwang JH, Yamada A, Ito K, Osawa S,
Connecting atomistic and experimental estimates of ideal strength. Kuramoto S, Suzuki N, Chen R, Saito T (2003) Development of
Phys Rev B Condens Matter Mater Phys 65:134111, doi:10.1103/ multi-functional titanium-alloy, “GUM METAL.” Ti-2003 Sci-
PhysRevB.65.134111 ence and Technology: Proceedings of the 10th World Conference
5. Furuta T, Kuramoto S, Hwang J, Nishino K, Saito T (2005) on Titanium: 1519–1526
Elastic deformation behavior of multi-functional Ti–Nb–Ta–Zr–0 10. Shan ZW, Mishra RK, Syed Asif SA, Warren OL, Minor AM
alloys. Mater Trans 46:3001–3007, doi:10.2320/matertrans. (2008) Mechanical annealing and source-limited deformation in
46.3001 submicrometre-diameter Ni crystals. Nat Mater 7:115–119,
6. Gutkin MY, Ishizaki T, Kuramoto S, Ovid’ko IA, Skiba NV doi:10.1038/nmat2085
(2008) Giant faults in deformed Gum Metal. Int J Plast 24:1333– 11. Shan ZW, Li J, Cheng YQ, Minor AM, Syed Asif SA, Warren OL,
1359 Ma E (2008) Plastic flow and failure resistance of metallic glass:
7. Withey E, Jin M, Minor A, Kuramoto S, Chrzan DC, Morris JW insight from in situ compression of nanopillars. Phys Rev B
(2008) The deformation of Gum Metal in nanoindentation. Mater Condens Matter Mater Phys 77:155419, doi:10.1103/PhysRevB.
Sci Eng A 493:26–32 77.155419

Potrebbero piacerti anche