You are on page 1of 6

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

net/publication/11063284

Ideal Pure Shear Strength of Aluminum and Copper

Article  in  Science · October 2002


DOI: 10.1126/science.1076652 · Source: PubMed

CITATIONS READS
465 1,132

3 authors, including:

Shigenobu Ogata Ju Li
Osaka University Oregon Health and Science University
256 PUBLICATIONS   3,418 CITATIONS    375 PUBLICATIONS   19,554 CITATIONS   

SEE PROFILE SEE PROFILE

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

Mechanics of energy materials View project

mechanics of nanowires View project

All content following this page was uploaded by Shigenobu Ogata on 03 May 2014.

The user has requested enhancement of the downloaded file.


REPORTS
have a much smaller size than the cooling 5. R. Maezono, S. lshihara, N. Nagaosa, Phys. Rev. B 58, room temperature to TP1, all the images are uniform,
11583 (1998). and from low temperature to TK2 during warming,
images at the same temperatures below TK1.
6. A. Asamitsu, Y. Moritomo, Y. Tomioka, T. Arima, Y. both the size and the contrast of the images are
This difference may be due to the way in Tokura, Nature 373, 407 (1995). nearly unchanged. To maintain the consistency be-
which the sample is cooled. The tip of the 7. C. Martin et al., Eur. Phys. J. B 16, 469 (2000). tween the resistivity hysteresis and the micromag-
MFM vibrates above the sample surface at 8. M. Uehara, S. Mori, C. H. Chen, S. W. Cheong, Nature netic hysteresis, we have simply reproduced the three
399, 560 (1999). highest temperature images in Fig. 1A and the three
the lever’s resonant frequency of 110 KHz, 9. M. B. Salamon, M. Jaime, Rev. Mod. Phys. 73, 583 (2001). lowest temperature images in Fig. 1C.
contacting the sample at the lowest point of 10. N. D. Mathur, P. B. Littlewood, Solid State Commun. 20. P. Levy et al., Phys. Rev. B 65, 140401 (2002).
each oscillation. At this contact point, the 119, 271 (2001). 21. L. Kong, S. Y. Chou, Appl. Phys. Lett. 70, 2043 (1997).
11. C. W. Yuan, E. Batalla, A. de Lozanne, M. Kirk, M. 22. F. Parisi, P. Levy, L. Ghivelder, G. Polla, D. Vega, Phys.
magnetic field applied by the tip is large (102 Tortonese, Appl. Phys. Lett. 65, 1308 (1994). Rev. B 63, 144419 (2001).
to 103 G) (21). Roughly, as the tip is scanned, 12. Q. Lu, C. Chen, A. de Lozanne, Science 276, 2006 23. A. Urushibara et al., Phys. Rev. B 51, 14103 (1995).
a strong periodic magnetic pulse with a fre- (1997). 24. Y.-A. Soh, G. Aeppli, N. D. Mathur, M. G. Blamire,
13. Materials and methods are available as supporting
㛬㛬㛬㛬
quency of 110 KHz is scanned over the sam- Phys. Rev. B 63, 020402 (2000).
material on Science Online. 25. , J. Appl. Phys. 87, 6743 (2000).
ple. We believe that upon cooling, although 14. Park Scientific Instruments (now Veeco Instruments), 26. M. Fäth et al., Science 285, 1540 (1999).
this scanning high-frequency localized mag- Santa Barbara, CA. 27. The work at Austin was supported by NSF grant
netic pulse may not change the relative FM 15. R. Desfeux, S. Bailleul, A. Da Costa, W. Prellier, A. M. DMA-0072834 and The Texas Center for Supercon-
Haghiri-Gosnet, Appl. Phys. Lett. 78, 3681 (2001). ductivity at the University of Houston. The work at
fraction (22), it can partially align the mag- 16. We recorded another set of cooling and warming Maryland was supported in part by NSF grant
netization of the domains and drive the mo- images using a temperature change rate of 0.3 K/min. MRSEC-00-80008. We thank S.-W. Cheong, N. D.
tion of the domain walls, leading to the merg- The temperature-dependent evolution of the FM do- Mathur, A. Guha, Q. Niu, Y. Tsui, and J. Lee for fruitful
mains is qualitatively the same as that of the sets discussions and T. Ruiter (Veeco Metrology Group)
ing and enlargement of the domains. On the shown in the text. We believe that our observation is for providing piezoresistive cantilevers.
other hand, the warming images were ob- largely due to the temperature change and not time
tained after the sample was cooled to the relaxation. Supporting Online Material
17. V. Podzorov, B. G. Kim, V. Kiryukhin, M. E. Gershen- www.sciencemag.org/cgi/content/full/1077346/DC1
lowest temperature in a magnetic field of 25 son, S.-W. Cheong, Phys. Rev. B 64, 140406 (2001). Materials and Methods
G. During this cooling process, the tip is far 18. A. Biswas et al., Phys. Rev. B 61, 9665 (2000). Movies S1 to S4
away from the sample. This cooling field 19. We emphasize that all of our sets of cooling and
warming images obtained on different areas of the 14 August 2002; accepted 9 September 2002
may not be strong enough to move the do- Published online 19 September 2002;
sample with different noncontact piezolevers show
main walls. When scanning during warming, the same temperature-dependent evolution of do- 10.1126/science.1077346
the domain walls are strongly pinned. As a mains. As described in the text, during cooling from Include this information when citing this paper.
result, the shape of the domains is unchanged.
As TP2 is approached, the average magneti-
zation decreases, which results in a rise in the
resistivity (Fig. 1B) (23). Our observations
Ideal Pure Shear Strength of
indicate that during cooling, the percolation
of the FM domains causes the steep resistiv-
Aluminum and Copper
ity drop, whereas during warming, the FM Shigenobu Ogata,1,2,3 Ju Li,1,4 Sidney Yip1*
conductive paths remain until near TP2, but
the decrease in the average magnetization
leads to the jump in resistivity. This also may Although aluminum has a smaller modulus in {111}具112៮ 典 shear than that of
explain why the knee in the resistivity is copper, we find by first-principles calculation that its ideal shear strength is
sharper during cooling than during warming. larger because of a more extended deformation range before softening. This
Below TK1 the enlargement of the domains fundamental behavior, along with an abnormally high intrinsic stacking fault
is mainly due to the merging of the domains, energy and a different orientation dependence on pressure hardening, are
and not to the increase of the FM volume traced to the directional nature of its bonding. By a comparative analysis of ion
fraction. Therefore, the resistivity changes relaxations and valence charge redistributions in aluminum and copper, we
slowly below TK1, as shown in Fig. 1B. arrive at contrasting descriptions of bonding characteristics in these two metals
Even well above TP2, there is still a slight but that can explain their relative strength and deformation behavior.
discernable contrast in some areas, which can be
more clearly seen in the supplemental movies. The minimum shear stress necessary to cause lations therefore would have important impli-
We propose that this is due to the magnetic permanent deformation in a material without cations for the understanding of the behavior
inhomogeneity above TP that is frequently ob- imperfections is fundamental to our concept of solids at the limit of structural stability.
served in similar CMR materials (24–26). of materials strength and its theoretical limits Results on stress-strain behavior of Al and Cu
Owing to the constraining effect of the under large strains (1, 2). With the possible in {111}具112៮ 典 shear, calculated with density
substrate, some effects observed in bulk sam- exception of recent nanoindentation measure- functional theory (DFT) and accounting for
ples may be suppressed or different for these ments (3), it has not been feasible to directly full atomic relaxation, have been reported
thin films (17). This might account for the measure the ideal shear strength of crystals. (4), where Cu was found to have a higher
sharper transitions in the thin films and the The demonstration that this property can be ideal shear strength than that of Al. Using
narrower hysteresis regions. This implies that reliably determined by first-principles calcu- various DFT methods and systematically
the temperature-dependent magnetic micro- cross-checking the results, we further inves-
structure in thin films may be modified due to 1
Department of Nuclear Engineering, Massachusetts tigated the shear strength and deformation of
the effect of the substrate. Institute of Technology, Cambridge, MA 02139, USA. Al and Cu and found instead that Al has the
2
Department of Mechanical Engineering and Systems, higher strength. Here, we report and substan-
3
Handai Frontier Research Center, Osaka University,
References and Notes Osaka 565– 0871, Japan. 4Department of Materials
tiate our findings by detailing the energetics
1. R. von Helmot, J. Wecker, B. Hopzapfel, L. Schulz, K. Science and Engineering, Ohio State University, Co- of shear deformation, the pressure-hardening
Samwer, Phys. Rev. Lett. 71, 2331 (1993). lumbus, OH 43210, USA. behavior, and valence charge redistribution
2. S. Jin et al., Science 264, 413 (1994).
3. C. Zener, Phys. Rev. 82, 403 (1951). *To whom correspondence should be addressed. E- during deformation. These considerations
4. A. J. Millis, Nature 392, 147 (1998). mail: syip@mit.edu show that the ideal shear strength and related

www.sciencemag.org SCIENCE VOL 298 25 OCTOBER 2002 807


REPORTS
properties such as stacking fault energies of constant). The unstable stacking energy ␥us, four other packages, Vienna ab Initio Simu-
Al and Cu can be accurately calculated and an important parameter in determining the lation Package (VASP) (7, 8), Cambridge
that the results can be rationalized by the ductility of the material (6), is ␥1(x0), where Serial Total Energy Package (CASTEP) (9),
underlying electronic structure. We suggest d␥1/dx(x0 ⬍ bp) ⫽ 0. It is instructive to WIEN2k (10), and ABINIT (11), with differ-
that bonding in Al is much more like a compare different ␥n(x) of the same slip sys- ent setups used to cross-check each other.
“hinged rod,” and we emphasize the impor- tem as n varies. The difference should be The results reported here are primarily those
tance of the breaking and reformation of di- relatively small from a local “glue” (shaded obtained from VASP with Perdew-Wang
rectional bonds as compared to the isotropic region in Fig. 1A) viewpoint where we take (PW91) generalized gradient approximation
“sphere-in-glue”–like behavior in Cu. the valence electron cloud to be the glue. We (GGA) exchange-correlation density func-
The intrinsic stacking fault energy, a mea- also have the asymptotic behavior at large n tional (12), ultrasoft (US) pseudopotential
sure of the energy penalty when two adjacent 2␥ twin共 x兲 (8), and Methfessel-Paxton smearing method
atomic planes in a crystal lattice are sheared ␥ n 共 x兲 ⫽ ␥ ⬁ 共 x兲 ⫹ ⫹ 共n⫺ 2 兲 (13) with 0.3-eV smearing width, and the cell
n
relative to each other, is known to play an being oriented as in Fig. 1A. The cutoff
important role in the structure and energetics (2) energies for the plane wave basis set for Al
of dislocations formed by slip processes. Al- where ␥twin(bp) is the unrelaxed twin and Cu are 162 and 292 eV, respectively.
though it is known experimentally that the boundary energy. The rate of convergence Table 1 shows the agreement of our results
intrinsic stacking fault energy is much larger to Eq. 2 reflects the localization range of with experimental and other calculations. To
in Al than in Cu, this finding has not been metallic bonding in a highly deformed bulk compute the equilibrium lattice constant a0,
related to their ideal shear strengths. For this environment. as well as relaxed and unrelaxed {111}具112៮ 典
purpose, we introduce a general function The DFT calculations, following the same shear moduli (Gr⬘ and Gu⬘, respectively) (4),
(Fig. 1A) procedure previously described (4), were per- we use a six-atom supercell of three {111}
E n 共 x兲 formed with our own plane wave code and layers. After relaxation, all stress components
␥ n 共 x兲 § , n ⫽ 1, 2, ... (1)
nS 0
where x is the relative displacement in the Table 1. Benchmark results, comparison of present calculations (Calc), experiments (Expt), and previous
slip direction between two adjacent atomic calculations (Oth calc). Dashes indicate that results are not available.
planes (we focus on {111}具112៮ 典 slip here),
Al Cu
En(x) is the increase in total energy relative to
Variable
its value at x ⫽ 0, with n ⫹ 1 being the Calc* Expt Oth calc Calc* Expt Oth calc
number of planes involved in the shearing
and S0 being the cross-sectional area at x ⫽ 0. a0 (Å) 4.04 4.03† 4.04‡ 3.64 3.62† 3.64§
The series of functions ␥1(x), ␥2(x), . . ., Gr⬘ (GPa) 25.4 27.4㛳 19 –25¶ 31.0 33.3㛳 26 –34¶
␥⬁(x) may be called the multiplane general- Gu⬘ (GPa) 25.4 27.6㛳 24 –30¶ 40.9 44.4㛳 36 – 44¶
␥sf (mJ/m2 ) 158 166# 143**, 39 45# (49)‡‡
ized stacking fault energy, with ␥1(x) being 164††
the conventional generalized stacking fault ␥us (mJ/m2 ) 175 — 183**, 158 — (210)‡‡
energy (GSF) (5) and ␥⬁(x) being the affine 224††
strain energy. The intrinsic stacking fault en-
*VASP, US-GGA, 18 ⫻ 25 ⫻ 11 Monkhorst-Pack kជ points. †(25) Al at temperature T ⫽ 0 K, Cu at T ⫽ 298 K.
ergy ␥sf is ␥1(bp), where bជ p ⫽ [112៮ ]a0/6 is the ‡(26) GGA. §(27) Full-potential linearized augmented plane wave method ( WIEN97 program), GGA. 㛳(28) calculat-
partial Burgers vector (a0, equilibrium lattice ed from elastic constants at T ⫽ 0 K. ¶(4) LDA. #(29). **(30) LDA. ††(31) LDA. ‡‡(5) LDA, unrelaxed.

Fig. 1. (A) Multiplane


generalized stacking
fault energy: n ⫽ 1, 2,
. . . , ⬁. (B) Pure shear
stress–displacement re-
sponses of Al (solid
squares) and Cu (open
squares) and (C) ion re-
laxation patterns in Al
and Cu. (D) Simple
shear stress–displace-
ment curves d␥⬁(x)/dx
(squares) compared to
d␥1(x)/dx (circles) in Al
(solid symbols) and Cu
(open symbols).

808 25 OCTOBER 2002 VOL 298 SCIENCE www.sciencemag.org


REPORTS
other than ␴13 (⫽␴31) are reduced to ⬍0.1 the scalar-relativistic approximation, respec- pand and contract in the y and x directions,
GPa. For the intrinsic and unstable stacking tively (10). Furthermore, the full-potential respectively.
energies ␥sf and ␥us, respectively, we use a projector augmented-wave (PAW) method The difference in relaxation patterns has
24-atom supercell of 12 layers for Al (10 (16) and the local density approximation important implications for the shear strength–
layers for Cu) with layers 1 and 12 facing (LDA) options of VASP also have been in- hardening behavior (Table 3), which also has
vacuum and shearing between layers 6 and 7. cluded in the cross-check. The difference be- been noted and discussed in terms of third-order
Relaxation of all layers along the ⬍111⬎ tween US and PAW is ⬍5% in the maximum elastic constant (18). When pressurized in the
direction is terminated when the force on stress values, and LDA consistently gives a 具110典 direction, Cu hardens, whereas Al softens
each atom is ⬍0.01 eV/Å. Al has a much stress that is 10 to 20% higher than that of substantially; however, if pressurized in the
larger ␥sf value than that of Cu, yet their ␥us GGA. With all methods used, Al is found to 具111典 direction, Al hardens substantially,
values are quite close. have ideal simple shear and pure shear whereas Cu softens slightly. These results show
For affine deformation calculations, we strengths that are higher than those of Cu that the pressure-hardening effect is highly de-
consider pure shear (␴ij ⫽ 0, except ␴13) and (Table 2) (17). pendent on orientation. A rough estimate of the
simple shear (x ⫽ 0 with no relaxations). The At equilibrium, Cu is considerably stiffer stress state at the displacement burst observed
corresponding stress-displacement curves are than Al; its bulk, simple, and pure shear in nanoindentation experiments (3) shows that
shown in Fig. 1, B and D, respectively. The (along {111}具112៮ 典) moduli are greater than the pressure components are at the level indi-
stress values are obtained from analytical ex- those of Al by 80, 65, and 25%, respectively. cated in Table 3. Thus, a large effect on the
pressions; they have been checked against However, Al has an ideal pure shear strength shear strength is to be expected. However, be-
numerical energy derivatives at several val- that is 32% larger than that of Cu because it cause the actual stress state is a complicated
ues of strain. After analyzing the effects of has a longer range of elastic strain before triaxial condition and given that the pressure-
smearing width, energy cutoff, and Brillouin softening (Fig. 1B): xmax/bp ⫽ 0.28 or hardening behavior is very anisotropic, one
zone integration kជ-point convergence, we es- ␥max ⫽ 0.20 in Al, which are the displace- cannot ascertain its real effect without an accu-
timate that the maximum stress values in ment and the engineering shear strain at the rate stress analysis. For this purpose, we apply
Table 2 have an uncertainty of ⬍0.1 GPa maximum shear stress, respectively, versus a method combining atomistic and finite-ele-
within each method used. ABINIT uses the xmax/bp ⫽ 0.19 or ␥max ⫽ 0.13 in Cu. The ion ment calculations (19).
Perdew-Burke-Ernzerhof GGA functional relaxations in these two metals are different Because Al has no core d states, its par-
(14) and norm-conserving Troullier-Martins (Fig. 1C). In Al, when the top atom slides tially occupied valence d bands are abnormal-
(TM) pseudopotential (15); CASTEP uses over the bottom atoms, the top atom hops in ly low in energy, which gives rise to direc-
PW91-GGA/US; our own plane wave code the z direction, and the bottom atoms contract tional bonding. At the six-atom interstice in
uses PW91-GGA/TM; and WIEN2k is a full- in the y direction (relaxation in x is almost Al, the pocket of charge density has cubic
potential augmented plane wave plus local zero). In Cu, there is almost no relaxation in symmetry and is very angular in shape, with
orbitals method, where the core and valence the z direction; the top atom translates essen- a volume comparable to the pocket centered
states are treated by the Dirac equation and tially horizontally, and the bottom atoms ex- on every ion (Fig. 2A). In Cu, there is no such

Fig. 2. Charge-density isosurface in (A) Al and (B) Cu and ⌬[␳(r, x)Vcell(x)]


(compared to a perfect crystal) along path a-b-c (␰, normalized path length
variable) during pure shear in (C) Al and (D) Cu. The figure shows
box-shaped extra charges in interstice volumes and their active evolutions
in Al, but not in Cu. The r max arrows point to positions of maximum
␳(r)Vcell along a-b-c at x ⫽ 0, indicating the size of the “atomic spheres”
centered at a, b, and c.

www.sciencemag.org SCIENCE VOL 298 25 OCTOBER 2002 809


REPORTS
Table 2. Ideal simple shear and pure shear strengths (␴ and ␴ , respectively).
u r
Table 3. Maximum shear stress under external
loading. P, hydrostatic pressure; ␴yy and ␴zz, nor-
Al Cu mal stress in the y and z directions, respectively.
Code
(number of kជ points)
␴u (GPa) ␴r (GPa) ␴u (GPa) ␴r (GPa) External stress Al (GPa) Cu (GPa)

VASP (12 ⫻ 17 ⫻ 7) 3.67* 2.76* 3.42 2.16 P ⫽ 0 (GPa) 2.84 2.16


VASP (18 ⫻ 25 ⫻ 11) 3.73 2.84 3.44† 2.15† P ⫽ 10 (GPa) 4.49 2.46
VASP (21 ⫻ 28 ⫻ 12) — — 3.45† 2.15† P ⫽ 20 (GPa) 5.90 2.84
VASP (27 ⫻ 38 ⫻ 16) 3.71* 2.84* — — ␴yy ⫽ ⫺10 (GPa) 1.78 3.12
CASTEP (13 ⫻ 22 ⫻ 9) — — 3.22† 2.10† ␴yy ⫽ ⫺20 (GPa) 1.41 3.54
CASTEP (17 ⫻ 29 ⫻ 12) 3.73* 2.84* — — ␴zz ⫽ ⫺3 (GPa) 3.64 2.03
WIEN2k (38 ⫻ 33 ⫻ 38)‡ 3.83* 2.98* 3.61† 2.23† ␴zz ⫽ ␴yy ⫽ ⫺10 (GPa) 3.98 4.38
ABINIT (12 ⫻ 17 ⫻ 7) 3.68* 2.84§ — — ␴zz ⫽ ␴yy ⫽ ⫺20 (GPa) 5.26 6.52
ABINIT (18 ⫻ 25 ⫻ 11) 3.71* 2.90* — —
Own code (12 ⫻ 17 ⫻ 7) 3.73* 2.89* 3.48† 2.18†
*Cell dimension from VASP (18 ⫻ 25 ⫻ 11) calculations. †Cell dimension from VASP (12 ⫻ 17 ⫻ 7) calcula- the electrons can redistribute well, and the
tions. ‡ One-atom cell (atomic sphere radius Rmt ⫽ 2.2 bohr; plane wave cutoff KmaxRmt ⫽ 10; charge density system does not incur a large energy penalty.
Fourier expansion cutoff Gmax ⫽ 20 ryderg ).
1/2 §Cell volume was relaxed by using ABINIT. In this work, we exploit the connection
between the generalized stacking fault energy
interstice pocket, and the charge density is and less accommodating, as manifested in a and the stress-strain response to show that the
nearly spherical about each ion (Fig. 2B). larger intrinsic stacking fault energy, for abnormally high ideal shear strength and in-
Thus, Al has an inhomogeneous charge dis- example. trinsic stacking fault energy of Al have the
tribution in the interstitial region because of To quantify our interpretation, we return same electronic-structure origin (namely, di-
bond covalency (20) and directional bonding to the behavior of the multiplane generalized rectional bonding). On one hand, directional
(21), whereas Cu has relatively homogeneous stacking fault energies in the form of stress- bonds give rise to a relatively longer shear
distribution and little bond directionality. To displacement functions d␥1(x)/dx and d␥⬁(x)/ deformation range, which accounts for the
probe these bonding characteristics further, dx (Fig. 1D). First, we note that for Cu, larger ideal shear strength of Al in relation to
we look at how the valence charge density d␥1(x)/dx and d␥⬁(x)/dx are not very different that of Cu; on the other hand, once the exist-
␳(r, x)Vcell(x) varies along a path in cell, across the entire range of shear, so the local ing bonds are broken and new bonds are
a-b-c, during pure shear, as atom b moves glue picture is appropriate. The fact that the formed with unfavorable bond angles, the
away from its initial nearest neighbor atom a sliding of a layer is effectively decoupled electrons cannot readjust easily, resulting in
(at x ⫽ x1) and takes on a new nearest from that of adjacent layers indicates that an anomalous intrinsic stacking fault energy
neighbor atom c (at x ⫽ x2 ). The ⌬[␳(r, bonding in Cu has nearly no bond-angle de- for Al. Our findings are supported by the
x)Vcell(x)] patterns, with Vcell being the cell pendence. On the other hand, the same func- detailed behavior of the valence charge den-
volume, for Al and Cu again show substantial tions behave much more differently in Al, sity obtained from first-principles calcula-
contrast. In Al (Fig. 2C), the maximum especially when x ⬎ xmax, at which the gra- tions that have been systematically cross-
change occurs halfway between the two near- dient reaches a maximum. Even in the range checked. We studied the pressure dependence
est ions, which indicates that when atoms of x ⬍ xmax, the relative magnitudes of of the shear strength and have shown its very
change neighbors, the breaking and reforma- d␥1(x)/dx and d␥⬁(x)/dx are opposite in order anisotropic character. These results suggest
tion of directional bonding is an important in Al as compared to those in Cu, suggesting that conventional crystal plasticity notions
activity. There is little such activity in Cu a possibly different nature of bonding. Sec- such as a scalar or pressure-independent yield
(Fig. 2D). ⌬[␳(r, x)Vcell(x)] mainly reflects an ond, the value of xmax is almost identical criterion based on critical resolved shear
accommodation process, like soft spheres between d␥1(x)/dx and d␥⬁(x)/dx in both Al stress, although successful for macroscopic
squeezing past each other by distorting their and Cu, with Al having the larger xmax value, face-centered cubic metals, should be viewed
own shape. A similar attempt to connect implying that the longer range directional cautiously when interpreting nanoindentation
stacking fault energy with redistribution and bonding in Al could be a more general feature experiments (3). Contemporary empirical po-
topological properties of charge density was than being specific to the affine strain energy tentials (24) may be useful for providing a
made recently (22). ␥⬁(x). Third, we see that when x ⬎⬎ xmax and qualitative description of the nonlinear,
The charge-density behavior just discussed, the directional bonds in Al are broken (con- anisotropic stress distribution under the
along with the relaxation patterns seen in Fig. firmed by a depleted charge at the interstice nanoindenter and for ascertaining the likely
1C, suggest a hinged-rod model to describe the in Fig. 2C), d␥1(x)/dx in Al stays positive for site and character of the instability; the quan-
shear strength for Al, in contrast to the conven- an extended range, whereas d␥1(x)/dx in Cu titative importance of these results remains to
tional “muffin-tin” or sphere-in-glue model for becomes negative quickly. Thus, although Al be scrutinized by more accurate ab initio
Cu. It is reasonable to think that when the and Cu have approximately the same unstable calculations.
bonding is directional (rodlike), a longer range stacking energy (Table 1), we see that when
of deformation can be sustained before break- the displacement x reaches bp and the config- References and Notes
ing than when the bonding is spherically sym- uration becomes an intrinsic stacking fault, 1. J. Wang, J. Li, S. Yip, S. Phillpot, D. Wolf, Phys. Rev. B
52, 12627 (1995).
metric, because of different geometrical factors Cu has recovered most of its losses in the 2. J. W. Morris Jr., C. R. Krenn, Philos. Mag. A 80, 2827
of charge-density decay with bond length. In sense of a low value of ␥sf , whereas Al has (2000).
covalent systems like Si (23) and SiC, we ver- recovered very little as its ␥sf value remains 3. A. Gouldstone, H. J. Koh, K. Y. Zeng, A. E. Giannako-
close to the ␥us value. The implication is that poulos, S. Suresh, Acta Mater. 48, 2277 (2000).
ified that during shear, the bonds generally do 4. D. Roundy, C. R. Krenn, M. L. Cohen, J. W. Morris,
not break until the engineering shear strain when a directional bond is broken, it is more Phys. Rev. Lett. 82, 2713 (1999).
reaches 25 to 35%, which is substantially larger difficult for the electrons to readapt. In con- 5. J. A. Zimmerman, H. Gao, F. F. Abraham, Model.
than those of metallic systems. Conversely, trast, for sphere-in-glue–type systems, even if Simul. Mater. Sci. Eng. 8, 103 (2000).
6. J. R. Rice, G. E. Beltz, J. Mech. Phys. Solids 42, 333
when the bonds do break, a directionally bond- the bond angles are wrong, as long as the (1994).
ed system can be expected to be more frustrated volumes fit as in the intrinsic stacking fault, 7. G. Kresse, J. Hafner, Phys. Rev. B 47, RC558 (1993).

810 25 OCTOBER 2002 VOL 298 SCIENCE www.sciencemag.org


REPORTS
8. G. Kresse, J. Furthmüller, Phys. Rev. B 54, 11169 are maintained at a publicly accessible Web site 27. F. Jona, P. M. Marcus, Phys. Rev. B 63, 094113 (2001).
(1996). (available at http://asm.mit.edu/sogata/AlCu/). 28. Landolt-Börnstein III/29a, Low Frequency Properties
9. V. Milman et al., Int. J. Quantum Chem. 77, 895 18. C. R. Krenn, D. Roundy, J. W. Morris Jr., M. L. Cohen, of Dielectric Crystals Second and Higher Order Elastic
(2000). Mater. Sci. Eng. A 317, 44 (2001). Constants, D. F. Nelson, Ed. (Springer-Verlag, New
10. P. Blaha, K. Schwarz, G. Madsen, D. Kvasnicka, J. Luitz, 19. J. Li, K. J. Van Vliet, T. Zhu, S. Yip, S. Suresh, Nature York, 1992).
WIEN2k, An Augmented Plane Wave ⫹ Local Orbitals 418, 307 (2002). 29. J. P. Hirth, J. Lothe, Theory of Dislocations ( Wiley,
Program for Calculating Crystal Properties ( Tech- 20. P. J. Feibelman, Phys. Rev. Lett. 65, 729 (1990). New York, ed. 2, 1982).
nische Universität Wien, Vienna, 2001). 21. J. C. Grossman, A. Mizel, M. Côté, M. L. Cohen, S. G. 30. J. Hartford, B. von Sydow, G. Wahnström, B. I. Lund-
11. The ABINIT code is a common project of the Univer- Louie, Phys. Rev. B 60, 6343 (1999). qvist, Phys. Rev. B 58, 2487 (1998).
sité Catholique de Louvain (Louvain-la-Neuve, Bel- 22. N. Kioussis, M. Herbranson, E. Collins, M. E. Eberhart, 31. G. Lu, N. Kioussis, V. V. Bulatov, E. Kaxiras, Phys. Rev.
gium), Corning Incorporated (Corning, NY ), and oth- Phys. Rev. Lett. 88, 125501 (2002). B 62, 3099 (2000).
er contributors (available at www.abinit.org). 23. Y. Umeno, T. Kitamura, Mater. Sci. Eng. B 88, 79 32. We thank J. W. Morris Jr. for comments on the
12. J. P. Perdew et al., Phys. Rev. B 46, 6671 (1992). (2002). manuscript. S.O. thanks Y. Shibutani and H. Kitagawa
13. M. Methfessel, A. T. Paxton, Phys. Rev. B 40, 3616 24. Y. Mishin, M. J. Mehl, D. A. Papaconstantopoulos, A. F. for discussions and acknowledges support by a Mu-
(1989). Voter, J. D. Kress, Phys. Rev. B 63, 224106 (2001). rata-Kaigai-Ryugaku-Syogakukai fellowship. J.L. and
14. J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 25. Landolt-Börnstein III/14a, Structure Data of Elements S.Y. acknowledge support by Honda R&D; the Air
77, 3865 (1996). and Intermetallic Phases. Elements, Borides, Carbides, Force Office of Scientific Research; NSFKDI and
15. N. Troullier, J. L. Martins, Phys. Rev. B 43, 1993 Hybrides, K.-H. Hellwege, A. M. Hellwege, Eds. ITR initiatives; and Lawrence Livermore National
(1991). (Springer-Verlag, New York, 1988). Laboratory.
16. G. Kresse, D. Joubert, Phys. Rev. B 59, 1758 (1999). 26. C. Stampfl, C. G. Van de Walle, Phys. Rev. B 59, 5521
17. For verification, the input files for the different codes (1999). 26 July 2002; accepted 11 September 2002

The Effect of Size-Dependent gaseous Pb and a Pb atom in a nanoparticle as


a detailed function of particle size (Fig. 1).

Nanoparticle Energetics on We then develop a model for predicting par-


ticle size evolution based on modified bond
additivity that better approximates the calori-
Catalyst Sintering metric data. Kinetic models of sintering rates
based on our modified bond-additivity esti-
Charles T. Campbell, Stephen C. Parker, David E. Starr mate are compared to the W-J model, as well
as to heuristic models that have been devel-
Calorimetric measurements of metal adsorption energies directly provide the oped because of the inaccuracies of the W-J
energies of metal atoms in supported metal nanoparticles. As the metal cov- model. We use our modified bond-additivity
erage increases, the particles grow, revealing the dependence of this energy on model to predict the sintering of gold parti-
particle size, which is found to be much stronger than predicted with the usual cles on TiO2 and compare these model results
Gibbs-Thompson relation. It is shown that this knowledge is crucial to accu- to experimental data we obtained via temper-
rately model long-term sintering rates of metal nanoparticles in catalysts. ature-programmed low-energy ion scattering
(TP-LEIS).
Metal nanoclusters, dispersed across the sur- catalyst supports can make it difficult to de- We can convert our calorimetric data for
face of an oxide or other support, can be termine particle sizes via microscopic meth- Pb on MgO(100) (16) to energy versus par-
much more active and selective as catalysts ods as a function of temperature. In the ab- ticle size, because there is good evidence that
than can larger metal particles (1, 2). How- sence of direct measurements, a commonly Pb grows on this surface as nearly hemispher-
ever, metal nanoclusters invariably sinter used approach for estimating the dependence ical particles with roughly constant number
(form larger clusters) under reaction condi- of particle energy on size has been to use the density of ⬃8.1 ⫻ 1011 islands/cm2, after the
tions, especially in some very important tech- Gibbs-Thompson relation, which states that first few percent of a monolayer. Using this
nical catalysts (2–8). The development of the chemical potential ( partial molar free en- particle density, the measured Pb surface
supported metal nanoclusters that resist long- ergy) of a metal atom in a particle of radius R, concentration can be converted directly into
term sintering has been hampered by the lack ␮(R), differs from that in the bulk [␮(⬁)] by the average number of atoms per particle, and
of a kinetic model that accurately predicts then into the average hemispherical particle
long-term sintering based on short-term mea- ␮共R) – ␮(⬁) ⫽ 2␥⍀/R (1) radius (Fig. 1). As can be seen, the stability of
surements. Without such a model, every where ␥ is the surface free energy of the the metal atoms in a Pb particle (that is, their
promising new catalyst must be tested for the metal and ⍀ is the bulk metals volume per heat of adsorption, relative to gaseous Pb)
actual length of time it must resist sintering in atom (4, 12). The use of this relation is decreases dramatically as the radius decreas-
application (⬃1 year). implicit in all current atomistic models of es below a few nanometers (17). For com-
We show here that the inclusion of accu- sintering (13–15), starting with the pioneer- parison to these direct measurements of the
rate size dependence of particle energies in ing models of Wynblatt and Gjostein (W-J) effect of the metal particle size on the metal
kinetic models is crucial in this respect. Little (3, 4). atom’s energy, the predictions of the Gibbs-
is currently known experimentally about the In this report, we use our recent micro- Thompson relation [Eq. 1, taking ␥ ⫽ 58.6
energetics of atoms within metal nanopar- calorimetric measurements of the heat of ad- ␮J/cm2 for Pb (18)] are also plotted in Fig. 1
ticles, although the energy of gaseous Sn sorption (qad) of Pb onto MgO(100) (16) to as “constant ␥ model”. Here, we neglect en-
clusters as a function of size was measured show that the energy of a metal atom in a tropy differences (12), so that qad(R) –
recently (9), and theoretical calculations con- nanoparticle increases much more dramati- qad(⬁) ⫽ –[␮(R) – ␮(⬁)] ⫽ –2␥⍀/R, where
tinue to address this important issue (10, 11). cally with decreasing size than predicted by qad(R) is the differential molar heat of adsorp-
The direct determination of particle energies the Gibbs-Thompson relation. Because the tion of Pb at fixed radius R. As is seen, Eq. 1
via calorimetry has been developed relatively Pb immediately forms Pb nanoparticles upon severely overpredicts the stability of Pb in
recently. In addition, the roughness of real adsorption, and these grow in radius smooth- small Pb particles, by ⬃60 kJ/mol at 1 nm
ly with increasing coverage, this measured radius (19). This shows that the surface en-
Department of Chemistry, Box 351700, University of adsorption energy versus coverage directly ergy increases substantially as the radius de-
Washington, Seattle, WA 98195–1700, USA. provides the difference in energy between creases below ⬃3 nm, which could be ex-

www.sciencemag.org SCIENCE VOL 298 25 OCTOBER 2002 811


View publication stats

You might also like