www.fgks.org   »   [go: up one dir, main page]

Academia.eduAcademia.edu
Acta Materialia 54 (2006) 119–130 www.actamat-journals.com Composition evolution of nanoscale Al3Sc precipitates in an Al–Mg–Sc alloy: Experiments and computations Emmanuelle A. Marquis a a,b , David N. Seidman b,* , Mark Asta b, Christopher Woodward b,c Materials Physics Department, Sandia National Laboratories, 7011 East Avenue, MS 9161, Livermore, CA 94550, United States b Materials Science and Engineering Department, Northwestern University, Evanston, IL 60208-3108, United States c Materials and Manufacturing Directorate, Air Force Research Laboratory, Wright Patterson AFB, OH 45433, United States Received 7 May 2005; received in revised form 18 August 2005; accepted 22 August 2005 Available online 25 October 2005 Abstract Controlling the distribution of chemical constituents within complex, structurally heterogeneous systems represents one of the fundamental challenges of alloy design. We demonstrate how the combination of recent developments in sophisticated experimental high resolution characterization techniques and ab initio theoretical methods provide the basis for a detailed level of understanding of the microscopic factors governing compositional distributions in metallic alloys. In a study of the partitioning of Mg in two-phase ternary Al–Sc–Mg alloys by atom-probe tomography, we identify a large Mg concentration enhancement at the coherent a-Al/Al3Sc heterophase 2 interface with a relative Gibbsian interfacial excess of Mg with respect to Al and Sc, Crel Mg , equal to 1.9 ± 0.5 atom nm . The correspond2 rel ing calculated value of CMg is 1.2 atom nm . Theoretical ab initio investigations establish an equilibrium driving force for Mg interfacial segregation that is primarily chemical in nature and reflects the strength of the Mg–Sc interactions in an Al-rich alloy.  2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. Keywords: Atom-probe tomography; Ab initio calculations; Al3Sc precipitates; Mg segregation; Coherent heterophase interface 1. Introduction With continuing rapid increases in computing power, computational modeling is increasingly augmenting traditional empirical investigations in the design of technologically advanced structural materials. Most high-performance metallic alloys contain multiple alloying elements, whose interactions govern the formation of strengthening second-phases, partitioning behavior, and segregation at internal interfaces such as grain boundaries or matrix/precipitate heterophase interfaces. Since such compositional variations are a critical factor governing the mechanical properties, theoretical understanding of interatomic interactions in multicomponent alloys is highly desirable from the standpoint of designing such materials. * Corresponding author. Tel.: +1 8474914391/9252943287. E-mail addresses: d-seidman@northwestern.edu, emarqui@sandia.gov (D.N. Seidman). In the case of Al alloys, scandium contributes significantly to improving strength by forming nanoscale coherent Al3Sc precipitates [1,2]. Additions of ternary elements aim at improving mechanical properties and nanostructural stability. Transition metals, such as Ti and Zr tend to decrease the coarsening kinetics of the L12 phase [3,4] whereas Mg resides exclusively in the a-Al matrix in solid solution [1]. This paper focuses on the effects of Mg additions on Al3Sc precipitation in Al–Sc alloys, which is important for optimizing the mechanical properties of these multicomponent alloys [5], and is part of a comprehensive study of the room temperature and elevated temperature (573 K) creep behavior of Al(Sc) based alloys [3–10]. From a fundamental viewpoint, addition of Mg to the a-Al/Al3Sc system, where Mg is an oversized atom (12.08% by radius and 40.82 by volume [11]) in solid solution in the a-Al matrix, constitutes a simple and welldefined system for studying elemental partitioning and heterophase segregation. This research builds on our initial 1359-6454/$30.00  2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.actamat.2005.08.035 120 E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 results obtained by high-resolution electron microscopy (HREM), showing that Mg additions alter the morphology of Al3Sc precipitates, that is, the {1 0 0} and {1 1 0} facets disappear and the precipitates become spheroidal [6,7]. Asta et al. recently published first-principles calculations demonstrating significant Mg segregation at coherent {1 0 0} a-Al/Al3Sc coherent interfaces [12,13]. The present study demonstrates how the behavior of Mg in a-Al/Al3Sc alloys can be both measured at the subnanoscale level using atom-probe tomography (APT) and predicted from ab initio calculations, thereby providing detailed insight into the driving forces for phase partitioning and interfacial segregation in this model two-phase ternary aluminum alloy. 2. Experimental procedures 2.1. Specimen preparations A cast Al alloy with nominal composition 2.2 Mg–0.12 at.% Sc was annealed at 618 C in air for 24 h (to ensure uniformity of the Mg concentration throughout the material), quenched into cold water, and then aged in air at 300 C for times between 0.33 and 1040 h. APT tips were obtained by a two-step electropolishing procedure. The initial polishing solution of 30 vol.% nitric acid in methanol was followed by a solution of 2 vol.% perchloric acid in butoxyethanol, used for final polishing to produce a sharply pointed tip, with a radius of curvature of less than 100 nm (Fig. 1). Field-ion microscopy (FIM) analyses were performed at a background pressure of 105 Torr consisting of a mixture of 80% Ne and 20% He; APT analyses were carried out under ultrahigh vacuum conditions (1010 Torr) for pulsed field-evaporation, which was performed with a pulse fraction (pulse voltage/steady state dc voltage) of 20% and a pulse frequency of 1500 Hz. Specimens were maintained at temperatures below 30 K. After aging for 0.5 h at 300 C, the precipitate number density, 4 ± 2 · 1022 precipitate m3 [8], is sufficiently high to perform APT random-area analysis. The error bars stated correspond to one standard deviation. After aging for 1040 h at 300 C, the number density of Al3Sc precipitates decreases to about 3 ± 1 · 1021 precipitate m3, and random area APT analysis is no longer an efficient technique. Hence, FIM imaging was first performed to locate precipitates that are just commencing to intersect the surface of a tip before starting an analysis. Fig. 1. Transmission electron micrograph of an APT tip, illustrating the coherency strain contrast of the Al3Sc precipitates after aging at 300 C for 5 h. 2.2. Data analysis A typical mass-to-charge state (m/n) spectrum is displayed in Fig. 2. The three isotopes of Mg are doubly charged at 12, 12.5 and 13 a.m.u., with no hydride formation. The measured isotopic abundances are 78.4 ± 0.3% for 24Mg2+, 10.5 ± 0.3% for 25Mg2+ and 11.1 ± 0.3% for 26 Mg2+, which agree favorably with the handbook values of 79%, 10% and 11%. Scandium is also doubly charged with multiple hydrides, exhibiting a possible overlap between Sc and singly charged Mg at m/n values of 24, 25 and 26 a.m.u. In some cases, peaks at 24, 25 or 26 a.m.u. are detected, but the correct mass ratio of the isotope 24 Mg to the other two isotopes, 25Mg, and 26Mg, was not found; therefore the corresponding ions were considered to be Sc hydrides. Since one of the objectives of this study is the measurement of Mg concentrations, this choice implies possibly under-estimating the Mg concentrations, in particular in the proximity of the Al3Sc precipitates. Fig. 2. Example of mass-to-charge state ratio (m/n) spectra from a tip containing an Al3Sc precipitate. E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 Data visualization and analysis of data sets were performed using a software code, ADAM 1.5, which was developed at Northwestern specifically for analyzing APT data [14]. 3. Results In the as-quenched state, Mg and Sc appear homogeneously distributed although the standard statistical v2-test used [15], which compares solute concentration distributions with binomial distributions of a perfectly random solid solution, does not rule out whether or not Mg or Sc atoms are homogeneously distributed in the a-Al matrix. Precipitates are observed after aging for 0.33 h (Fig. 3(a)). A cluster search algorithm with a maximum separation distance of 0.7 nm is used to isolate the Sc atoms constituting the precipitates. Taking into account the spherical morphology of the precipitates [8], the composition was measured in spherical shells centered on the center-of-mass of the selected Sc atoms. Eight precipitates were analyzed yielding an average concentration of 22.4 ± 2.8 at.% Sc. The radius of these Al3Sc precipitates, estimated using the radius of gyration of the Sc atoms, is 121 between 0.8 and 1.4 nm. Magnesium atoms are also present inside the precipitates at a level, 4.3 at.%, which is approximately a factor of two greater than the average Mg concentration in the a-Al matrix. After further aging, the precipitate radius increases (Figs. 3(b) and (c)) and reaches about 4 nm at 1040 h (Fig. 4). The spatial resolution of APT is illustrated in Fig. 4, where an analysis performed near the 110 crystallographic pole reveals the {2 2 0} atomic planes perpendicular to the analysis direction. The Sc concentration of the precipitates increases after 0.5 h and remains constant thereafter within experimental error, xSc = 27.4 ± 1.5 at.% Sc (Table 1). The small discrepancy with the nominal composition, i.e. 25 at.%, is likely to be due to the different evaporation fields of Al and Sc that could not be accommodated despite the low tip temperature <30 K. On the other hand, the Mg concentration decreases with increasing aging time to 0.9 at.% after 1040 h aging. The Mg concentration within the Al3Sc precipitates is non-uniform with an enhancement at their centers as shown in the example of Fig. 5. For aging times longer than 2 h, a distinct Mg concentration enhancement at the a-Al/Al3Sc heterophase interface is also observed. The example displayed in Fig. 5 is a proximity histogram [16] that calculates the average composition in shells of 0.4 nm thickness at a given distance from the a-Al/Al3Sc heterophase interface. The interface is defined by an isoconcentration surface at 18 at.% Sc. The Mg concentration enhancement at the a-Al/Al3Sc heterophase interface is 200%, which is localized within 2 nm at this heterophase interface. The maximum Mg concentration at this heterophase interface decreases slightly during the early aging times and thereafter remains constant within experimental error. Fig. 6 displays the normalized average Mg concentration profile taken from the centers of the precipitates for each aging time. The profiles are temporally invariant after 2 h. 4. Discussion 4.1. Mg segregation: experimental measurement Fig. 3. Three-dimensional reconstruction of an analyzed volume, displaying only the Sc atoms, from a specimen aged at 300 C for: (a) 0.33 h; (b) 0.5 h; and (c) 5 h. Mg segregation at the a-Al/Al3Sc interface is constant with aging time, which suggests that the measured concentrations represent an equilibrium behavior. Moreover, the root-mean-square diffusion distance of Mg in Al is given pffiffiffiffiffiffiffiffi by 6Dt, where the factor 6 is for diffusion in 3-dimensions; D = 1.6 · 1016 m2 s1 is the diffusion coefficient of Mg in Al [17] at 300 C and t is the aging time; the values are about 1.3 lm after 0.5 h aging at 300 C and 60 lm after 1040 h. The precipitate spacing is evaluated using the square-lattice pffiffiffiffiffiffiffiffiffiffiffiffiffi spacing approximation, which is given by hRi ¼ 3 4p=3f , where ÆRæ is the mean precipitate radius and f = 0.53 vol.% is the calculated volume fraction of Al3Sc precipitates at 300 C. The precipitate spacing is 10 nm for ÆRæ equal to 1.1 nm (0.5 h aging), and 39 nm for ÆRæ equal to 4.2 nm (1040 h aging). The root-mean-square 122 E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 Fig. 4. Three-dimensional reconstruction of an Al3Sc precipitate after aging at 300 C for 1040 h; Al atoms are in blue, Mg atoms in green and Sc atoms in red. Table 1 Al3Sc precipitate composition as function of aging time at 300 C Time (h) Number of precipitates Al (at.%) Mg (at.%) Sc (at.%) As quenched 0.33 0.5 2 5 30 1040 0 8 4 10 3 2 15 – 73.3 ± 3.5 68.8 ± 5.2 71.2 ± 1.4 68.9 ± 2.1 71.0 ± 2.8 69.4 ± 2.8 – 4.3 ± 2.6 4.09 ± 1.5 2.3 ± 0.6 3.1 ± 1.1 2.5 ± 0.8 0.9 ± 0.3 – 22.4 ± 2.8 28.5 ± 1.4 26.5 ± 1.4 28.1 ± 2.5 26.5 ± 2.3 29.2 ± 2.4 Fig. 6. Mg concentration normalized by the average Mg concentration in the a-Al matrix for four different aging times (0.5, 2, 5, and 1040 h) at 300 C. 0 Crel Mg ¼ CMg  CSc Fig. 5. Proximity histogram of an Al3Sc precipitate after aging at 300 C for 1040 h; the colored areas correspond to the interfacial excesses of Al (blue), Mg (green) and Sc (red). diffusion distance of Mg is therefore always significantly greater than the average center-to-center precipitate spacing and the system is at the very least in local thermodynamic equilibrium with respect to Mg segregation at the a-Al/Al3Sc heterophase interface. The relative Gibbsian interfacial excess concentration of Mg with respect to Al and Sc provides a quantitative thermodynamic description of the observed equilibrium Mg segregation, and it is given by [18]: 0 xaAl xaMg  xaAl xaMg 0 0 xaAl xaSc  xaAl xaSc  CAl 0 0 0 0 xaMg xaSc  xaMg xaSc xaAl xaSc  xaAl xaSc ; ð1Þ where CMg, CSc and CAl are the Gibbsian interfacial ex0 cesses of Mg, Sc, and Al, respectively, and the xaj and xaj are the concentrations of component j (j = Al, Sc or Mg) in phase a (Al) and a 0 (Al3Sc). Fig. 5 displays the Gibbsian excess quantities for Al (negative value), Mg and Sc (positive values) as areas under the concentration curves in the proximity histogram [19]. The relative Gibbsian excesses of Mg with respect to Al and Sc are listed in Table 2 for the different aging times. The variations observed are within the experimental errors. The similar values of the Gibbsian excess of Mg relative to Al and Sc for all aging times implies that the system is at the very least in local thermodynamic equilibrium and the average value is Crel Mg ¼ 1:9  0:5 atom nm2 . For the Al–Sc–Mg system, the calculations of the solute concentrations at the interface of a growing Al3Sc precipitate predict a depletion of Mg at 123 E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 Table 2 Precipitate radius (from [7]), relative Gibbsian interfacial excess of Mg as function of aging time at 300 C, and maximum Mg enhancement factor Aging time (h) 0.5 2 5 30 1040 Radius (nm) a Crel Mg matrix cmax Mg =cMg 1 1.5 ± 0.7 2.8 ± 0.3 – 1.7 ± 0.4 2.7 ± 0.2 2 1.7 ± 0.3 2.0 ± 0.2 – 2.22 ± 0.45 2.3 ± 0.5 4.2 1.85 ± 0.38 1.9 ± 0.5 a Calculated using Eq. (1). the a-Al/Al3Sc interface (see below and Table 3). Thus our observation of a positive value of the relative Gibbsian interfacial excess of Mg with respect to Al and Sc is further evidence that we are observing a true thermodynamic equilibrium excess quantity. The decrease in interfacial free energy associated with Mg segregation can be estimated using [20]: ! oc rel CMg ¼  ; ð2Þ olMg T;P Crel Mg where is the relative Gibbsian excess of Mg with respect to Al and Sc, and lMg is the chemical potential of Mg. Assuming an ideal solid-solution, the chemical potential is lMg ¼ l0 þ k B T lnðxaMg Þ, and it yields the following expression:   Crel oc Mg k B T ¼ . ð3Þ oxMg T;P xaMg At 300 C, the measured relative Gibbsian excess of Mg 2 with respect to Al and Sc is Crel Mg ¼ 1:9  0:5 atom nm , and the average Mg concentration in the matrix is xaMg ¼ 2:2  0:3 at.% Mg. Assuming Crel Mg varies linearly with concentration, the integration of Eq. (3) for xaMg varying from 0 to 0.022 yields a decrease in the interfacial free energy of 15 mJ m2. Several previous APT studies reported segregation behavior of solute atoms at partially semi-coherent or semi-coherent heterophase interfaces [22,23]. In the present study, the coherency state of the a-Al/Al3Sc interface is known from HREM observations, and was determined to be perfectly coherent for all cases presented. Coherency loss may occur when the precipitate diameter is sufficiently large. For the lattice parameter misfit at 300 C, d  0.62%, between the Al matrix containing 2.2 at.% Mg and the Al3Sc phase, the spacing between the misfit dislocations is a/d, where a  0.2 nm is the spacing between {2 0 0} Table 3 Values of solute concentrations at the precipitate/matrix interface calculated from Eqs. (4) and (5) for spherical a 0 (Al3Sc) precipitates of radius R growing in a supersaturated a (Al) matrix in a ternary Al–Sc–Mg alloy 0 planes; this yields a critical precipitate diameter for loss of coherency of approximately 30 nm, which is much greater than the precipitate diameter measured after aging at 300 C for less than 1040 h. The exact shape of the precipitates observed by APT needs to be carefully considered. Firstly, the observations may be subject to experimental artifacts, such as asymmetry of the tip, or misalignment of the analysis direction. Also, the results depend strongly on the field evaporation behaviors of the a-Al matrix and Al3Sc precipitates, which may be significantly different. The measured local atomic density of the a-Al matrix is indeed higher than that of the precipitate phase, with an experimental density ratio equal to about 1.4. It is consistent with the bright imaging of the Al3Sc precipitates in FIM mode, indicating that the precipitates exhibit a small protrusion effect, due to their higher evaporation field. The width of the heterophase interface may therefore be explained by the artificially higher magnification of the Al3Sc precipitates and possibly ion trajectory effects [24]. The approximately spheroidal shape of the Al3Sc precipitates observed by HREM is, however, reproduced and no significant distortions between the lateral dimension and the depth dimension are observed. 4.2. Driving force for Mg segregation In this section we employ first-principles calculations to analyze the microscopic factors governing the pronounced enhancement of Mg at the Al/Al3Sc interface. In this analysis we first make use of the theoretical framework provided by a model of diffusion-limited precipitate growth kinetics, to ascertain whether the measured Mg enhancement may be due simply to capillary effects. The analysis leads to the conclusion that the Mg enhancement measured by APT cannot be interpreted simply as reflecting the effects of capillarity and solute-flux balance in a model of diffusion-limited growth. We subsequently use first-principles calculations to establish that the interfacial enhancement corresponds to an equilibrium segregation phenomenon driven by a chemical driving force reflecting the attractive interactions between Mg and Sc solute atoms. 0 R (nm) ^xaSc (at. %) ^xaMg (at.%) ^xaSc (at.%) ^xaMg (at.%) 2 4 1 2.5 · 104 1.8 · 104 9.0 · 105 2.2 2.2 2.4 25 25 25 3.4 · 107 4.0 · 107 4.8 · 107 These values were derived assuming values for the interfacial free energy, r equal to 0.175 J m2 [21] that are independent of composition and radius, as discussed in the text. 4.2.1. Interface solute concentrations in a model of diffusion-limited growth In this section we explore whether the solute–concentration profiles measured by APT can be interpreted as representing steady-state solutions to the solute diffusion equation, subject to the boundary conditions imposed by 124 E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 local thermodynamic equilibrium and flux balance at the growing precipitate/matrix interface. In this analysis we employ a model for diffusion-limited growth of precipitates in a ternary alloy due to Kuehmann and Voorhees [25]. While in a binary alloy the condition of local thermodynamic equilibrium uniquely determines the interfacial solute compositions at the interface during diffusion-limited growth, the situation is qualitatively different in a ternary alloy. As described in [25], the conditions of local thermodynamic equilibrium provide only three equations (equality of chemical potentials for each of the chemical species), which do not determine uniquely the four independent values of the interface solute concentrations (two in each of the two phases). Within a diffusion-limited-growth model for a ternary alloy, the additional constraint required to fix the interface compositions arises from the requirement of mass conservation (flux balance) at the moving precipitate/matrix interface. As a consequence, interface compositions in a ternary alloy are dictated not only by bulk and interfacial thermodynamic properties, but also by the ratio of the diffusivities in the matrix phase, which affect the relative fluxes of each solute species. For spherical precipitate geometries, employing a meanfield, quasi-steady-state solution to the diffusion equation (neglecting off-diagonal terms in the diffusion matrix), interface compositions can be derived from the following equations [25]: 0 0 2c  V i ði ¼ Al; Sc; MgÞ; R ð^xaSc  x1 Sc Þ ; ð^xaMg  x1 Mg Þ 0 lai ð^xaMg ; ^xaSc Þ  lai ð^xaMg ; ^xaSc Þ ¼ a0 ð^xSc  ^xaSc Þ DSc ¼ a0 a D ð^xMg  ^xMg Þ Mg 0 0 ð4Þ ð5Þ where ^xaSc , ^xaMg , ^xaSc , and ^xaMg denote mole fractions of Sc and Mg on the Al (a) and Al3Sc (a 0 ) sides of the interface for a precipitate of radius R, c is the interfacial free energy, the 0 variables lai and lai correspond to bulk chemical potentials, and V i is the partial molar volume for species i in 1 the precipitate (a 0 ) phase. The variables x1 Sc and xMg correspond to far-field solute concentrations in the matrix phase. The first three equations represented by Eq. (4) correspond to the well-known Gibbs–Thomson conditions incorporating the effect of capillarity in the formulation of the conditions for local thermodynamic equilibrium [26], while Eq. (5) reflects the constraint imposed by solute flux balance. Since values for the chemical potentials in the ternary Al3Sc intermetallic phase, required in the solution of Eqs. (4) and (5), are unavailable from experimental measurements, we have employed first-principles free-energy models in calculations of the equilibrium phase compositions. The bulk free energies were derived within a model of non-interacting substitutional defects through a generalization of the approach outlined in [13,29–31]. Such a noninteracting defect model for the free energy is expected to be highly accurate for the dilute solute concentrations considered in the present study. In this approach to modeling the thermodynamics of the bulk a- and a 0 -phases, the freeenergy models take the following form: F a ¼ F a0 þ k B TxAl ln xAl þ xSc ½DF aSc þ k B T ln xSc  þ xMg ½DF aMg þ k B T ln xMg ; 3n 0 0 a0 ;Al Al Al k B TxAl þ k B T ln xAl F a ¼ F a0 þ Al ln xAl þ xSc ½DF Sc Sc  4 o a0 ;Al Al þxAl Mg ½DF Mg þ k B T ln xMg  1n a0 ;Sc Sc Sc Sc k B TxSc þ Sc ln xSc þ xAl ½DF Al þ k B T ln xAl  4 o a0 ;Sc Sc þxSc ½DF þ k T ln x  ; B Mg Mg Mg ð6Þ where in the a-Al phase xAl, xSc and xMg are the mole fractions of Al, Sc and Mg, while in the a 0 -Al3Sc phase xAl Al and xSc Al denote Al mole fractions on the Al and Sc sublattices, Sc Al Sc respectively, and similarly for xAl Sc , xSc , xMg and xMg . The a quantities DF j for the a-Al phase in Eq. (6) denote the free energy to form a substitutional impurity of type j in pure Al; at zero temperature these are simply the heats of solution (DEaj ), while at finite temperature DF aj generally contains entropic contributions of vibrational and electronic origins. 0 Similarly, for the a 0 -Al3Sc phase DF aj ;k denotes the free energy to form substitutional impurity j on sublattice k. The various defect energies entering in Eq. (6) were computed from first-principles using the ab initio total-energy and molecular-dynamics program VASP (Vienna ab initio simulation package) developed at the Institut für Materialphysik of the Universität Wien [32–34]. In these calculations, use was made of ultrasoft pseudopotentials [35], the local-density approximation, and an expansion of the electronic wave functions in plane waves. Further details of the calculations can be found in [31]. Calculations of the point-defect energies entering Eq. (6) were performed employing 64-atom supercells, and were derived from the relation DEj/;k ¼ Ej/;k  E/ þ ðEk  Ej Þ, where Ej/;k is the energy of the supercell of phase / containing an impurity of type j on site k, E/ is the energy of the supercell for stoichiometric phase /, and Ej and Ek denote the energies per atom of pure species j and k in their respective equilibrium crystal structures, respectively. The results of the supercell defect-energy calculations are presented in Fig. 7. It is seen that Sc is calculated to have a large and negative heat of solution in a-Al, while the corresponding value for Mg is near zero. In the a 0 -Al3Sc phase, all substitutional defects are seen to have relatively large and positive formation energies. The DE values given in Fig. 7 were used in Eq. (6) to derive bulk chemical potentials, which formed the basis for the computation of the equilibrium concentrations given in Table 3. In these calculations we also included vibrational-entropy 0 contributions to F a0 , F a0 and DF aSc , since these contributions have been shown to be crucial for reproducing the measured binary Al–Al3Sc solvus boundary compositions [29]. The details of the vibrational-entropy calculations are given in [29,30]. E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 125 Fig. 7. Calculated substitutional point-defect energies in Al and Al3Sc phases. Results for a-Al are plotted on the right and those for a 0 -Al3Sc on the left. Solid and hatched bars denote defect energies on the Sc and Al sublattices in the a 0 -Al3Sc phase, while the white bars denote heats of solution for Mg and Sc in a-Al. Table 3 lists values of the interface concentrations calculated from Eqs. (4) and (5) for precipitates of radii 2 and 4 nm at an aging temperature of 300 C. For comparison we also list compositions corresponding to two-phase equilibrium between bulk (R ! 1) a and a 0 -phases in a ternary alloy with the composition Al–2.2 at.% Mg–0.12 at.% Sc considered experimentally. For this alloy composition, the bulk equilibrium compositions derived from first-principles (final row of Table 3) are in very reasonable agreement with the values of ^xaSc ¼ 7:2  104 at.% and ^xaMg ¼ 2:2 at.%, derived independently by Murray [36] from empirical free-energy models. In our calculations of the interface compositions listed in Table 3, values for the far-field matrix compositions in Eq. (5) were taken as x1 Sc ¼ 0:014  0:001 at.% and x1 ¼ 2:35  0:02 at.%, as measured by APT in alloys aged Mg for 1040 h [8]. We further made use of the following measured values for the solute diffusivities in Al: DSc = 8.84 · 1020 m2 s1 and DMg = 1.62 · 1016 m2 s1 [17,32]. Additionally, we employed a value for c equal to 0.175 J m2, as derived from first-principles calculations for planar interfaces between pure Al and stoichiometric Al3Sc [21] (c is assumed to be constant and independent of composition). A comparison of the results in the first two rows of Table 3 with those corresponding to bulk phases (final row) shows that the effects of capillarity and solute flux balance are estimated to give rise to relatively small (10%) changes in the matrix Mg concentration at the growing precipitate/matrix interface. Thus, the pronounced interfacial enhancement of Mg measured by APT cannot be interpreted simply as reflecting the effects of capillarity and solute flux balance in a model of diffusion-limited precipitate growth. Additional first-principles calculations, discussed in the following section, suggest instead that the observed interfacial segregation of Mg reflects an equilibrium segregation effect, arising from electronic interactions between Sc and Mg atoms in Al. 4.2.2. Equilibrium segregation energies for Mg at a coherent Al/Al3Sc interface First-principles VASP calculations were conducted to investigate the energetics of Mg solute atoms in the vicinity of planar coherent Al/Al3Sc interfaces aligned parallel to {2 0 0} crystallographic planes. The heats of solution for Mg solute atoms were computed as a function of distance in the vicinity of a planar Al/Al3Sc interface employing supercell geometries with periodic boundary conditions. Preliminary results from these calculations were reported in [12] and further details were given in [13]. The supercells used in these calculations contained periodic vectors in the plane of the interface (a1 and a2) with lengths four times the face-centered cubic (fcc) nearest-neighbor spacing: a1 = a(2, 2, 0), a2 = a(2, 2, 0), where a is the fcc lattice constant. In the z-direction, normal to the interface, the periodic length corresponded to 14 fcc unit cell dimensions. The total number of atoms in these supercells was 412. The results presented below were derived using a value for the in-plane lattice constant (a) equal to that for the bulk Al3Sc phase; the dimension of the supercell in the z-direction was then adjusted to give zero total zz-stress for a pure Al/Al3Sc interface (without Mg additions). Additional calculations were performed using a value of the in-plane lattice constant equal to that for bulk fcc Al, and only minor differences in the calculated segregation energies were obtained. In this section we report the solution energy of a single Mg atom placed at various crystal sites in the vicinity of the Al/Al3Sc (0 0 2) interface. For calculations in which Mg substitutes on the Al side of the interface, the supercells contained a total of nine unit cells of fcc Al and five unit cells of Al3Sc. Similarly, for calculations of Mg impurity formation energies on the Al3Sc side of the interface we employed supercells with five fcc-Al unit cells and nine Al3Sc unit cells. Magnesium impurity formation energies were derived for substitution both on Al and Sc sites in the supercell; in all cases the formation energies on Sc lattice positions were sufficiently high to yield negligible equilibrium Mg concentrations on these sites. 126 E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 In Fig. 8 we plot the formation energies (DE) for substituting Mg for Al as a function of distance across the coherent (0 0 2) Al/Al3Sc planar interface. The values of DE calculated on the Al3Sc side of the interface are roughly 0.6 eV larger than in pure Al, indicative of the strong energetic preference for the partitioning of Mg to the matrix phase. In terms of the APT measurements, the most important feature of the results shown in Fig. 8 is the significantly negative formation energy calculated on the Al side of the interface at the crystallographic site corresponding to a second neighbor of the interface Sc atoms (the site labeled ‘‘1’’ in Fig. 8). The value of DE at this site is computed to be 0.1 eV lower than the heat of solution for Mg atoms in bulk Al. This segregation energy (DE) thus provides a reasonably strong driving force for the equilibrium segregation of Mg atoms to the (0 0 2) coherent Al/Al3Sc interface. To make explicit contact with the experimental results, the energies derived from the first-principles calculations have been used within a mean-field (Bragg–Williams) model for the configurational free energy (similar to Eq. (6)) to compute the equilibrium solute composition profiles across a planar Al/Al3Sc {2 0 0} interface at 300 C. Within this approximation, the equilibrium concentration of Mg atoms at site j, xMg(j), is given as xMg ðjÞ ¼ x0Mg expf½DEðjÞ DEð1Þ=k B T g, where x0Mg is the bulk Mg concentration in a-Al, DE(j) is the formation energy of a substitutional Mg impurity at site j near the interface, and DE(1) is the heat of solution in bulk a-Al (far from the interface). The calculated Mg concentration profile is shown in Fig. 9, and was derived assuming a bulk Mg concentration in Al of DE(1) = 2.4 at.%. The plot features a five-to sixfold enhancement of the xMg in the plane positioned one lattice constant from the interface Sc atoms. In comparison to the … APT results for Mg plotted in Fig. 5, the calculated concentration profile is considerably narrower with a larger value for the Mg enhancement factor. These differences between the experimental and theoretically calculated composition profiles may be due to the effects discussed in Section 4.1 associated with the APT technique. Specifically, the measured width of the concentration profile at the Al/Al3Sc interface is probably greater than the real width due to effects associated with the field-evaporation process [24], which make a concentration profile appear broader than it is (see Section 4.1). Due to these differences in the width, a more meaningful comparison between experiment and theory can be made in terms of the integrated area under the Mg concentration profiles. Specifically, the relative Gibbsian interfacial adsorption segregation of Mg with respect to Al and Sc provides a quantitative thermodynamic description of the degree of equilibrium Mg segregation, If we interpret the Mg segregation to reflect an equilibrium segregation effect, we can estimate the values of the Gibbsian excess quantities for Al (negative value), Mg and Sc (both positive values) as the areas under the concentration curves in the proximity histogram as displayed in Fig. 5 [29]. The value thus derived 2 from the experimental data is Crel Mg = 1.9 ± 0.5 atom nm , which is found to be independent of aging time after 0.5 h at 300 C. By comparison, Crel Mg derived from the calculated composition profiles is 1.2 atom nm2. Thus, experiments and calculations yield values for the relative Gibbsian interfacial excess of Mg, which are approximately the same within the experimental uncertainty of ± 0.5 atom nm2. The good level of agreement between experiment and theory thus supports strongly the conclusion that the measured interfacial enhancement of Mg reflects a pronounced 2 4 1 3 … 0.6 Al ∆ E (eV) 0.4 0.2 Al3Sc 0 -0.2 -5 -4 -3 -2 -1 z (Lattice Constants) 0 1 Fig. 8. Calculated formation energies for substitution of Mg for Al as a function of distance from the (0 0 2) a-Al/Al3Sc interface. A projection of the interface atomic structure of the interfacial region is shown on top, with white and gray circles denote Al positions in planes parallel to the interface, and black circles corresponding to Sc sites. In this projection the large circles denote atomic sites positioned 0.5 of a lattice constant below those indicated by the smaller circles. The formation energies are plotted as a function of distance normal to the interface in the lower plot; white and gray circles denote values of DE derived for the corresponding sites in each plane shown in the projection above. 127 E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 Fig. 9. Calculated Mg concentration profile near the a-Al/Al3Sc interface. Compositions have been derived from the results in Fig. 8 assuming a bulk Mg concentration in the Al phase of 2.4 atomic %. Sc-Sc 200 Interaction Energy(meV) equilibrium effect associated with the segregation of this species to planar coherent Al/Al3Sc heterophase interfaces. It is noteworthy that the calculated segregation energy derived from the first-principles supercell calculations are found to be highly insensitive to the imposed value of the in-plane lattice parameter (the value of which was varied in the calculations between the equilibrium lattice constants for pure Al and Al3Sc phases as described above). Furthermore, the calculated segregation energy is found to change by less than 10% if the interface positions are fixed at ideal fcc positions. That is, the calculated segregation energy is found to be insensitive to the state of strain at the interface. Thus the origin of the calculated segregation energy appears not to originate from elastic strain energy. Rather, we interpret the result as being a manifestation of the nature of Mg–Sc electronic interactions in an Al-rich alloy. In support of this interpretation we have computed the solute interaction energies up to fourth neighbor distances for Mg–Mg, Sc–Sc and Mg–Sc pairs in pure Al employing VASP and 216-atom supercells (6 · 6 · 6 primitive fcc unit cells). The results are shown in Fig. 10. Strong and relatively long-ranged interactions are derived for both Mg– Sc and Sc–Sc pairs. The magnitude and oscillating nature of the Sc–Sc results are consistent with theoretical models for transition-metal interactions in Al due to Carlsson and Moriarty [38,39]. For both Sc–Sc and Mg–Sc we obtain repulsive interactions at first and third neighbors, and attractive interactions at second and fourth. The magnitude and spatial variation of these interactions were found to be primarily electronic in origin; very similar values to those plotted in Fig. 10 were obtained in supercell calculations where the atoms were constrained to their bulk fcc lattice sites (i.e., removing any elastically induced contribution to the interactions). In light of the results in Fig. 10, the overall magnitude of the calculated segregation energy, as well as the preferred binding site for Mg at the Al/Al3Sc {2 0 0} heterophase interface, can be rationalized as follows. The supercell calculations yield a preferred binding site for Mg, labeled ‘‘1’’ in Fig. 8, that contains the maximum 100 Mg-Mg 0 Sc-Mg -100 -200 1.0 1.2 1.4 1.6 1.8 2.0 r/rnn Fig. 10. Mg–Mg, Mg–Sc and Sc–Sc interaction energies in an Al host, calculated as a function of separation r normalized by the nearestneighbor distance (rNN). number of both second neighbors (one per Mg atom) and fourth neighbors (four per Mg atom) to interface Sc atoms, while featuring no nearest or third-nearest neighbor repulsive Mg–Sc pairs. Assuming that the alloy energetics are dominated by pair interactions, the interaction energies plotted in Fig. 10 can be used to derive a magnitude for the Mg binding energy at site ‘‘1’’ equal to 0.14 eV. This result is in very good overall agreement with the value derived from direct supercell calculations (see Fig. 8). Furthermore, the Mg impurity energy at the site labeled ‘‘2’’ in Fig. 8 is computed to be near zero in the direct supercell calculations. This result is again consistent with a bond-counting analysis considering that this site features only third neighbor interactions with interface Sc atoms; the magnitude of the third-neighbor Mg–Sc interactions are shown to be relatively small in Fig. 10. Finally, both sites ‘‘3’’ and ‘‘4’’ in Fig. 8 are calculated to have relatively large positive Mg impurity formation energies on the order of 0.2 eV. This result is again consistent with bond 128 E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 counting, given the fact that these sites have two nearestneighbor bonds with interface-Sc atoms, each contributing approximately 0.1 eV of interaction energy to DE according to Fig. 10. Overall, the very good agreement between the directly calculated formation energies at the interface sites, and the corresponding values derived from a relatively simple bond-counting analysis, suggests strongly that Mg segregation at the Al/Al3Sc interface can be interpreted as being a reflection of strong Mg–Sc electronic interactions in Al. The type of bond-counting analysis described in the previous paragraph provides a convenient framework for analyzing the magnitude of the anisotropy associated with Mg segregation to a coherent Al/Al3Sc interface. Such an analysis is interesting in light of the experimental observations by Marquis et al. [8,14] showing that Mg additions to Al– Sc lead to a pronounced change in the morphology of Al3Sc precipitates from highly faceted (with well-developed {1 0 0}, {1 1 0} and {1 1 1} facets) to a spheroidal morphology. Due to the higher areal densities of sites containing attractive second and fourth neighbor interactions with interface Sc atoms, Mg segregation is estimated to be roughly a factor of three and 2.5 larger for {1 1 0} and {1 1 1} interfaces, respectively, relative to the lower-energy {1 0 0} orientation. We consider the analysis for {1 0 0} versus {1 1 0} orientations in detail in Fig. 11. This figure highlights with gray and hatched circles the atomic sites that are predicted to have appreciable binding energies for Mg. In the case of {1 0 0} discussed, there is one attractive site per area a2, with a segregation energy of approximately 0.1 eV. By comparison, the pffiffi{1 ffi 1 0} interface contains two attractive sites per area, 2a2 . The site colored gray in Fig. 11 contains two second-nearest-neighbor and five fourth-neighbor Sc atoms. From the interaction energies in Fig. 10, this site is thus predicted to have a binding energy of approximately 0.2 eV, leading to a site concentration at 300 C that is predicted to be nearly pure Mg (assuming a bulk composition of 2.4 at.% and using a Bragg–Williams model as above). The second binding site for a {1 1 0} interface, indicated by the hatched circle in Fig. 11, shares one fourth neighbor with interface-Sc atoms and thus has a relatively small binding energy of roughly 0.025 eV, leading to a concentration enhancement of about 1.7 times the bulk concentration at 300 C. The net result is an enhancement of Crel Mg by approximately a factor of three for a {1 1 0} interface relative to {1 0 0}. A similar analysis leads to the prediction of a value for Crel Mg for {1 1 1} orientations that is enhanced by approximately a factor of two relative to {1 0 0}. Using the Gibbs adsorption theorem, Mg segregation is predicted to lower the free energy of {1 0 0}, {1 1 1} and {1 1 0} interfaces by approximately 10, 20 and 30 mJ m2, respectively. Magnesium segregation thus lowers the free energy of {1 1 0} and {1 1 1} orientations by 10–20 mJ m2 relative to {1 0 0} orientations. In [13], calculated interfacial free energies for Al/Al3Sc interfaces in binary Al–Sc showed a {1 0 0} vs. {1 1 1} anisotropy of 25 mJ m2 at 300 C (with the free energy of {1 1 0} estimated to be similar to {1 1 1} ). The present results thus suggest that Mg segregation leads to a substantial reduction in the anisotropy of the Al/Al3Sc interfacial free energies, consistent with the observed reduction in precipitate faceting induced by the addition of Mg [8,13] and the resulting spheroidal morphology. Fig. 11. The projections compare the atomic geometries near {1 0 0} and {1 1 0} interfaces. As in Fig. 8, small and large circles denote atomic sites separated by 0.5 of a lattice spacing in the direction out of the page. Black and white circles denote Sc and Al sites, respectively. The solid and hatched gray circles denote Al sites near the interface displaying attractive binding energies for Mg, as described in the text. where xMg and xSc are the average concentrations of Mg and Sc atoms, and gbMg–Sc is an average Gibbs binding free energy between Mg and Sc atoms for first and second nearest-neighbors. In the case of a random solid solution, the binding energy is zero therefore leading to an expected 4.3. Heterogeneous nucleation of Al3Sc precipitates? As the aging time increases from 2 to 1040 h, the Mg segregation peak observed at early time splits into two peaks, one at the interface separated from a peak at the center of the precipitates (Fig. 6). In particular after 1040 h aging, 15–42 Mg atoms were detected at the center of the Al3Sc precipitates. It is speculated that Al3Sc precipitation occurs as a result of interactions between Mg and Sc atoms with vacancies, which leads to a faster nucleation rate than in the binary Al–Sc alloy, as demonstrated by the microhardness measurements [5]. Indeed, from an analysis of the early stages of decomposition, the experimental number of Mg– Sc dimers in the reconstruction is somewhat greater than the expected number of pairs in the case of a perfectly random solid-solution. The experimental number of Mg–Sc dimers defined by a maximum separation distance between the Mg and Sc atoms of 0.5 nm, is 1082 for a total number of ions, N, equal to 1.154 · 106; this yields a concentration of Mg–Sc dimers equal to 0.0937 at.%. The theoretical concentration of dimers (first nearest and second nearest-neighbors), xMg–Sc, defined by the number of dimers divided by the total number of atoms, is given by: ! gbMg–Sc xMg–Sc ¼ 18xMg xSc exp ; ð7Þ kB T E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 number of first and second nearest-neighbor dimers, given by 18xMgxSc = 0.0693 at.%, The measured concentration of dimers is 60% of the actual value, assuming an ion detection efficiency of 60%. An estimate of the average Gibbs binding free energy in the as-quenched state is then calculated to be 0.040 eV. After 20 min aging, the measured concentration of dimers is 0.0617 at.%, compared to the random solution concentration of 0.0484 at.%, which yields an average binding energy of 0.037 eV. A positive value of gbMg–Sc indicates an attractive interaction between atoms, in agreement with the first principles calculations in Fig. 10. Al3Sc precipitation occurs as a result of interactions between Mg and Sc atoms with vacancies, which leads to a faster nucleation rate than in the binary Al–Sc alloy. The high concentration of quenched-in vacancies and their high mobility, even at room-temperature, could explain a fairly high number density of Mg–Sc dimers and, during the formation of a stable nucleus involving the Mg–Sc dimers, Mg atoms may get trapped within the nanoscale growing Al3Sc precipitates, as Sc diffuses to the precipitates. We were unable to find diffusion data for any element in Al3Sc in the literature. The presence of a Mg-rich Al3Sc precipitate core after a long aging time, i.e. 1040 h, (Fig. 7) indicates, however, a very small diffusivity of Mg diffusivity of Mgffi in the Al3Sc phase. An estimate of the qffiffiffiffiffiffiffiffi ffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Al3 Sc in Al3Sc at 300 C is obtained using hd 2 i ¼ 6DMg t, where d is the precipitate diameter and the factor of six is from three-dimensional diffusion. The estimated diffusivity is therefore given by: 3 Sc DAl ffi Mg hd 2 i . 6t ð8Þ For t = 1040 h, Eq. (8) yields a value of 2 · 1023 m2 s1, which is approximately seven orders of magnitude smaller than the diffusivity of Mg in Al at 300 C (1.62 · 1016 m2 s1) [37]. This indicates that it is difficult for Mg to diffuse through the Al3Sc phase to the matrix, which explains the shape of the Mg concentration profiles observed in Figs. 6 and 7, even though the equilibrium ternary Al–Mg–Sc phase diagram [36] does not predict any Mg solubility in the Al3Sc phase. And the first principles calculations presented in Section 4.2 also indicate essentially no solubility of Mg in Al3Sc. 5. Conclusions  Magnesium segregation occurring at the perfectly coherent a-Al/Al3Sc heterophase interface was studied experimentally using atom-probe tomography (APT) for an Al–2.2 Mg–0.12 at.% Sc alloy aged at 300 C.  The thermodynamic equilibrium Mg segregation behavior corresponds to a measured relative Gibbsian excess of Mg with respect to Al and Sc of 1.9 ± 0.5 atoms nm2. 129  In addition to the Mg segregation at perfectly coherent aAl/Al3Sc heterophase interfaces we also detected Mg at the centers of Al3Sc precipitates, Fig. 7, which is kinetically trapped since Mg is insoluble in Al3Sc. The diffusivity of Mg in Al3Sc is estimated to be 2 · 1023 m2 s1, which is approximately seven orders of magnitude smaller than the diffusivity of Mg in Al at 300 C (1.62 · 1016 m2 s1).  The effects of capillarity and solute flux balance are theoretically estimated to give rise to relatively small (10%) changes in the matrix Mg concentration at the growing coherent a-Al/Al3Sc heterophase interface. Therefore, the pronounced interfacial enhancement of Mg measured by APT cannot be interpreted simply as reflecting the effects of capillarity and solute flux balance in a model of diffusion-limited precipitate growth.  The segregation of Mg at the coherent a-Al/Al3Sc heterophase interface was studied theoretically employing ab initio calculations. These calculations demonstrate that the driving force for segregation of Mg is due to electronic interactions rather than elastic strain relaxation associated with highly over-sized Mg atoms. The calculated value of the relative Gibbsian excess of Mg with respect to Al and Sc is ca. 1.2 atoms nm2, which is in good agreement with the experimental value.  The results of the supercell defect-energy calculations are presented in Fig. 8. It is seen that Sc is calculated to have a large and negative heat of solution in a-Al, while the corresponding value for Mg is near zero. In the a 0 -Al3Sc phase, all substitutional point defects are seen to have relatively large and positive formation energies.  We have computed the solute interaction energies up to fourth neighbor distances for Mg–Mg, Sc–Sc and Mg– Sc pairs in pure Al employing VASP and 216-atom supercells (6 · 6 · 6 primitive fcc unit cells), see results in Fig. 11. Strong and relatively long-ranged interactions are derived for both Mg–Sc and Sc–Sc pairs. The magnitude and oscillating nature of the Sc–Sc results are consistent with theoretical models for transitionmetal interactions in Al.  Using the Gibbs adsorption theorem, Mg segregation is predicted to lower the free energy of {1 0 0}, {1 1 1} and {1 1 0} heterophase interfaces by roughly 10, 20 and 30 mJ m2, respectively. Magnesium segregation thus lowers the free energy of {1 1 0} and {1 1 1} orientations by 10–20 mJ m2 relative to {1 0 0} orientations.  These present theoretical results suggest that Mg segregation leads to a substantial reduction in the anisotropy of the a-Al/Al3Sc interfacial free energies, which is consistent with our HREM observations of a reduction in precipitate faceting induced by the addition of Mg to Al–Sc alloys [8]. Acknowledgments This research is supported by the United States Department of Energy, Basic Sciences Division, under contracts 130 E.A. Marquis et al. / Acta Materialia 54 (2006) 119–130 DE-FG02-98ER45721 (EAM and DNS) and DE-FG0201ER45910 (MDA) and the Air Force Research Laboratory, the Air Force Office of Scientific Research under contract F33615-01-C-5214 (CW). The authors thank Professors P.W. Voorhees and D.C. Dunand for very interesting discussions, Dr. D. Isheim for help in using an atomprobe tomograph, and Alcoa Inc. and Ashurst Inc. for supplying the Al–Sc master alloy. The calculations performed in this work made use of resources at the National Energy Research Scientific Computing Center under Contract No. DE-AC03-76SF00098, and the DOD High Performance Computer Modernization Program at the Aeronautical Systems Center-Major Shared Resource Center on the IBM-SP3. References [1] Toporova LS, Eskin DG, Kharakterova ML, Dobatkina TB. Advanced aluminum alloys containing scandium. Amsterdam: Gordon & Breach; 1998. [2] Royset J, Ryum N. Int Mater Rev 2005;50:2. [3] Fuller CB, Murrey JL, Seidman DN. Acta Mater 2005;53:5401; Fuller CB, Seidman DN. Acta Mater 2005;53:5415. [4] van Dalen ME, Dunand DC, Seidman DN. MS&T 2003 – Affordable metal matrix composites for high performance applications II. Pittsburgh, PA: TMS; 2003. p. 195. [5] Marquis EA, Seidman DN, Dunand DC. Acta Mater 2003;51:4751. [6] Marquis EA. Ph.D. thesis, Northwestern University, 2002. [7] Marquis EA, Seidman DN. Acta Mater 2001;49:1909. [8] Marquis EA, Seidman DN. Acta Mater 2005;53:4259. [9] Marquis EA, Dunand DC. Scripta Mater 2002;47:503. [10] Fuller CB, Seidman DN, Dunand DC. Acta Mater 2003;4803:51. [11] King HW. J Mater Sci 1966;1:79. [12] Asta M, Ozolins V, Woodward C. JOM 2001;53(9):16. [13] Marquis EA, Seidman DN, Asta M, Woodward C, Ozolins V. Phys Rev Lett 2003;91:036101. [14] Hellman OC, Vandenbroucke J, Blatz du Rivage J, Seidman DN. Mater Sci Eng A 2001;327:29. [15] Miller MK, Smith GDW. Atom probe microanalysis. Pittsburgh: MRS; 1989. [16] Hellman OC, Vandenbroucke JA, Rüsing J, Isheim D, Seidman DN. Microsc Microanal 2000;6:437. [17] Rothman SJ, Peterson NL, Nowicki LJ, Robinson LC. Phys Status Solidi B 1974;63:K29. [18] Dregia SA, Wynblatt P. Acta Metall Mater 1991;39:771. [19] Krakauer BW, Seidman DN. Phys Rev B 1993;48:6724. [20] Lupis CHP. Chemical thermodynamics of materials. Englewood Cliffs, NJ: Prentice Hall; 1983. [21] Asta M, Foiles SM, Quong AA. Phys Rev B 1998;57:11265. [22] Isheim D, Csencsits R, Seidman DN. Mater Res Soc Symp Proc 2001;652:Y1021. [23] Isheim D, Seidman DN. Mater Metall Trans A 2002;33:2317. [24] Vurpillot F, Bostel A, Blavette D. Appl Phys Lett 2000;76:3127. [25] Kuehman CJ, Voorhees PW. Metall Trans A 1996;27:937. [26] The precipitates considered in this work are coherent, and the bulk a and a 0 phases are known to have a lattice parameter mismatch of roughly 0.6% in the ternary Al–Sc–Mg alloy at the aging temperature of 300 C. As a consequence, the equations governing thermodynamic equilibrium, Eq. (1), should generally be modified to account for the elastic energy associated with coherency strains (e.g., [27]). For the precipitate radii relevant to the Al–Sc–Mg system (2–4 nm), the elastic energy density derived from EshelbyÕs model [28] of a spherical misfitting inclusion is found to be more than an order of magnitude smaller than the capillarity term 2c/R. We have therefore neglected detailed consideration of the effects associated with coherency strains in the analyses of interface solute concentrations. [27] Larché FC, Cahn JW. Acta Metall 1978;53:26. [28] Eshelby JD. Proc Roy Soc London Ser A 1957;241:376. [29] Ozolins V, Asta M. Phys Rev Lett 2001;86:448. [30] Asta M, Ozolins V. Phys Rev B 2001;64:094104. [31] Woodward C, Asta M, Kresse G, Hafner J. Phys Rev B 2001;63:094103. [32] Kresse G, Hafner J. Phys Rev B 1993;47:558. [33] Kresse G, Hafner J. Phys Rev B 1994;49:14251. [34] Kresse G, Furthmuller J. Computat Mater Sci 1996;6:15. [35] Kresse G, Furthmuller J. Phys Rev B 1996;54:11169. [36] Murray JL, private communication. [37] Fujikawa SI. Defect Diff Forum 1997;143–147:115. [38] Zou J, Carlsson AE. Phys Rev B 1993;47:2961. [39] Moriarty JA, Widom M. Phys Rev B 1997;56:7905.