22
Pergamon J. Pkys. Cfim. SoMa Vol. 55, No. to, pp. 10354057, 1994 Copyright Q 1994 Eiswier Science Ltd Printed in Great Britain. All tights nserwd 0022~3697/94 $7.00 + 0.00 ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT GRAIN BOU~ARIES: EXPERIMENTS AND SIMULATIONS D. N. SEIDMAN, B. W. KRAKAUER AND D. UDLER Department of Materials Science and Engineering and Materials Research Center, R. R. McCormick School of Engineering and Applied Science, Northwestern University, Evanston, IL 602083108, U.S.A. Abstract-This paper presents two atomic scale approaches to study grain boundary (GB) segregation phenomena. The first is an experimental one that combines ~an~ssion electron microscopy (TEM) with atom-probe field-ion microscopy (APFIM~~FIM~~M-to measure quantitatively the Gibbsian interfacial excess of solute at GBs whose five macroscopic degrees of freedom are first measured by TEM; with this approach it is possible to explore systematically GB phase space. APFIM is also used to determine segregation profiles with atomic resolution. An application is presented for this combined experimental approach for a single phase Fe(Si) alloy. The second involves Monte Carlo simulations of soluteatom segregation at GBs in bicrystals of single-phase f.c.c. alloys; this approach is also used to systematically explore GB phase space. The atoms are allowed to interact via long-range continuous embedded atom method potentials, and so-called transmutational ensemble is employed. The results show that, unlike the previously investigated Au-Pt system, the (002) twist boundaries are enhanced in solute atoms on both sides of the phase diagram. For low-angle (002) twist boundaries on the Pt-rich side the atomic sites enhanced in solute con~ntration are arranged in hourly-like structures centered on the square grid of primary grain boundary dislocations. While for the same boundaries on the Ni-rich side the atomic sites enhanced in solute concentration are located in bipyramidal regions based on the squares cells of the same grain boundary dislocations. Thus, the atomic sites that are enhanced on one side of the phase diagram are not affected on the other side and vice versa. Keywords: A. interfaces, A. metals, C. electron microscopy, D. defects, D. microstructure. 1. ~ODU~IO~ AND GENERAL BACKGROUND The ultimate extent to which we can predetermine the properties of materials depends on our ability to control and tailor the properties of internal interfaces and surfaces [l-6]. This ability has resulted in a field called interfacial engineering [7-91, whose foun- dations are based on a fundamental knowledge of the phases, structure, chemists and kinetics of internal interfaces. Interfacial engineering is a crucial com- ponent of materials design, particularly for tailoring the mechanical properties of structural materials- fracture mode, toughness, diffusional creep and creep cavitation embrittlement-that are frequently deter- mined by phenomena occurring at internal interfaces under applied external stresses. More generally, pro- cesses occurring in the interfacial volume affect many important physical IlO, 111 and technological proper- This festschrift article is written on the occassion of Bob Ball&i’s 70th birthday in recognition of his many seminal contributions to materials science. David Seidman thanks Bob Balluffi for stimulating his interest in atomic scale observations and for support and friendship since his graduate studies at the University of Illinois at Urbana- Champaign. ties of a wide range of bulk materials; these processes include phase transitions, short-circuit diffusion [12, 131,nucleation of new bulk phases, nucleation of voids in electromigration damage of metal intercon- nects, etc. These phenomena are sensitive to inter- facial phases, structure and chemistry. The interfacial region where two grains of a single- phase material meet is called a grain boundary (GB) or a homophase interface and similarly the region where two grains of two differing phases meet is denoted a heterophase interface or interphase inter- face. In the last 25 years a significant body of experimental and theoretical research has focused on the structures of homophase interfaces [14]. This work has emphasized both the dislocation structures and the positions of atoms at internal interfaces using transmission-el~tron (TEM) and high-resolution electron microscopies (HREM) as well as X-ray diffraction techniques [15-171. There have also been efforts to correlate the structures and local chem- istries of internal interfaces by employing energy dispersive X-ray and electron energy loss spectro- xopies in conjunction with transmission electron or scanning transmission electron microscopy [18]. 1035

ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Pergamon

J. Pkys. Cfim. SoMa Vol. 55, No. to, pp. 10354057, 1994 Copyright Q 1994 Eiswier Science Ltd

Printed in Great Britain. All tights nserwd 0022~3697/94 $7.00 + 0.00

ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT GRAIN BOU~ARIES:

EXPERIMENTS AND SIMULATIONS

D. N. SEIDMAN, B. W. KRAKAUER AND D. UDLER Department of Materials Science and Engineering and Materials Research Center, R. R. McCormick School of Engineering and Applied Science, Northwestern University, Evanston, IL 602083108, U.S.A.

Abstract-This paper presents two atomic scale approaches to study grain boundary (GB) segregation phenomena. The first is an experimental one that combines ~an~ssion electron microscopy (TEM) with atom-probe field-ion microscopy (APFIM~~FIM~~M-to measure quantitatively the Gibbsian interfacial excess of solute at GBs whose five macroscopic degrees of freedom are first measured by TEM; with this approach it is possible to explore systematically GB phase space. APFIM is also used to determine segregation profiles with atomic resolution. An application is presented for this combined experimental approach for a single phase Fe(Si) alloy. The second involves Monte Carlo simulations of soluteatom segregation at GBs in bicrystals of single-phase f.c.c. alloys; this approach is also used to systematically explore GB phase space. The atoms are allowed to interact via long-range continuous embedded atom method potentials, and so-called transmutational ensemble is employed. The results show that, unlike the previously investigated Au-Pt system, the (002) twist boundaries are enhanced in solute atoms on both sides of the phase diagram. For low-angle (002) twist boundaries on the Pt-rich side the atomic sites enhanced in solute con~ntration are arranged in hourly-like structures centered on the square grid of primary grain boundary dislocations. While for the same boundaries on the Ni-rich side the atomic sites enhanced in solute concentration are located in bipyramidal regions based on the squares cells of the same grain boundary dislocations. Thus, the atomic sites that are enhanced on one side of the phase diagram are not affected on the other side and vice versa.

Keywords: A. interfaces, A. metals, C. electron microscopy, D. defects, D. microstructure.

1. ~ODU~IO~ AND GENERAL BACKGROUND

The ultimate extent to which we can predetermine the

properties of materials depends on our ability to

control and tailor the properties of internal interfaces

and surfaces [l-6]. This ability has resulted in a field

called interfacial engineering [7-91, whose foun-

dations are based on a fundamental knowledge of the

phases, structure, chemists and kinetics of internal

interfaces. Interfacial engineering is a crucial com-

ponent of materials design, particularly for tailoring

the mechanical properties of structural materials-

fracture mode, toughness, diffusional creep and creep

cavitation embrittlement-that are frequently deter-

mined by phenomena occurring at internal interfaces

under applied external stresses. More generally, pro-

cesses occurring in the interfacial volume affect many

important physical IlO, 111 and technological proper-

This festschrift article is written on the occassion of Bob Ball&i’s 70th birthday in recognition of his many seminal contributions to materials science. David Seidman thanks Bob Balluffi for stimulating his interest in atomic scale observations and for support and friendship since his graduate studies at the University of Illinois at Urbana- Champaign.

ties of a wide range of bulk materials; these processes include phase transitions, short-circuit diffusion [12, 131, nucleation of new bulk phases, nucleation of voids in electromigration damage of metal intercon- nects, etc. These phenomena are sensitive to inter- facial phases, structure and chemistry.

The interfacial region where two grains of a single- phase material meet is called a grain boundary (GB) or a homophase interface and similarly the region where two grains of two differing phases meet is denoted a heterophase interface or interphase inter- face. In the last 25 years a significant body of experimental and theoretical research has focused on the structures of homophase interfaces [14]. This work has emphasized both the dislocation structures and the positions of atoms at internal interfaces using transmission-el~tron (TEM) and high-resolution electron microscopies (HREM) as well as X-ray diffraction techniques [15-171. There have also been efforts to correlate the structures and local chem- istries of internal interfaces by employing energy dispersive X-ray and electron energy loss spectro- xopies in conjunction with transmission electron or scanning transmission electron microscopy [18].

1035

Page 2: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1036 D. N. SEIDMAN er al.

Interfacial segregation and phase transitions occur on an atomic scale, thus if one is to compare experimen- tal results with computer simulations [19-221 one must use a probe with atomic scale resolution to determine the chemical identities and positions of individual atoms. Before addressing specific chal- lenges, problems, and techniques toward this end, we consider general background material on the thermo- dynamics of interfaces, interface phases and GB phase transitions. This material provides a frame- work for our experimental and computer simulation studies.

Gibbs developed the concepts of the thermodyn- amics of interfaces [23] and defined the basic exten- sive thermodynamic properties. Cahn [24] extended Gibbs’ analysis to include the five macroscopic de- grees of freedom (DOFs) that define the geometric character of a GB; he assumed that the three micro- scopic DOFs are relaxed to their equilibrium values. The five macroscopic DOFs are defined when a GB is “created” by taking a unit vector, c, in one of the two grains with direction cosines, cr and cl, and then rotating grain 1 with respect to grain 2 about c by an angle, 0. The plane of the interface is defined by a unit normal vector, II. It is assumed that the microscopic DOFs of a GB are relaxed to values that minimize the total free energy of a bicrystal. Cahn’s treatment results in a generalized expression for the interfacial free energy per unit area of interface, 0, or more simply the interfacial free energy:

da = da(T, P, pi, strain, c,, c,, 0, n,, n,), (1)

where T, P and pi are the temperature, hydrostatic pressure, and chemical potentials of each of the

components. The strain in eqn (1) refers to nonhydro- static stresses and their concomitant strains [25]. Thus, da is a function of many variables and, there- fore, the phase space of an interface is in general a

priori unknown and quite complex; each point in the multidimensional phase space represents a thermo- dynamic state. The exploration of this phase space is ideally achieved in a systematic manner by moving along a trajectory of one variable with all the remain- ing variables held fixed. We have performed both experiments and computer simulations utilizing this approach for the Gibbsian interfacial excess of solute (F,) in b.c.c. and f.c.c. alloys, respectively [26,27]. The quantity I’, is the basic measure of interfacial segregation; for example, for a GB in a single bulk phase containing only two components this quantity is:

F2 = -(aa/apc,)T, P, spectjied DOFs; (2)

where F2 is the excess of the solute, 2, in an interfacial layer over what is present in a reference system containing the same amount of solvent. To explore a GB’s phase space properly and to obtain a physically meaningful value of F2, careful attention must be paid to the five macroscopic DOFs-c, II, 8.

The quantity r, plays a central role, for example, in the Bic+Thomson-Wang theory of the effects of segregation on fracture [28]. The phenomenological picture of intergranular fracture is a transition be- tween two limiting states: (a) a bicrystal and (b) two totally separated crystals with two new free surfaces. In a thermodynamic model this transition occurs reversibly under applied external forces that work against the forces of cohesion 129-311. In the presence of a segregating solute species, the change in the free energy of the system due to GB decohesion can be written as:

(3)

where a, and ub are the free energies of the two new surfaces, and Agb and Ag, are the Gibbs free energies of segregation at a GB and a free surface, respect- ively. It follows from eqn (3) that for GB fracture the work of decohesion, 2ui,,, depends on both the crystallographic plane (II) along which the fracture occurs and on the misorientation of the two grains (0 and c or Z). (The value of the quantity Z is the inverse of degree of coincidence of the two grains comprising a bicrystal in three dimensions; for example, Z = 5 implies one in five sites is in coincidence between the two three-dimensional grains.) Furthermore, eqn (3) demonstrates that a solute that segregates more strongly to the free surfaces than the GB embrittles that boundary, for example, Bi in Cu; whereas a species that segregates more strongly to a GB than to the free surfaces ductilizes that boundary, for example, B in N&Al [32]. Thus, the behavior of TZ in GB phase space plays an important role in ultimately designing ductile materials.

It is important to recall that a GB contains many types of sites with local environments that differ appreciably from those of the bulk lattice; those sites may either attract or repel solute atoms. Thus, the distribution of segregated solute atoms at a GB can be heterogeneous on an atomic scale. Solute-atom segregation may be examined experimentally at an atomic scale employing atom-probe field-ion mi- croscopy (APFIM), or alternatively, via various com- puter simulation techniques, for example, Monte Carlo (MC) or molecular dynamics (MD) [33,34]. Simulations play an important role in studying segre- gation as they sample the C + 6 dimensional phase space in a systematic manner; our efforts have con-

Page 3: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Solute-atom segregation at grain boundaries 1037

centrated on studying f.c.c. binary alloys for which reliable embedded atom method (EAM) potentials are available [35]. For example, we have performed detailed MC studies of solute segregation in (002) twist boundaries, for dilute alloys in the Au-P1 and Ni-Pt systems for 0” < 6 ~45” (by symmetry 0” < 0 < 360”). The picture of the atomistics of segre- gation obtained via MC simulations presents an extraordinary experimental challenge [36,371. Exper- imentally, we have measured directly via APFIM, for example, Fsi , in a -Fe as a function of c, n, 8. These are the first absolute measurements of Fz for GBs with a known structure in any alloy [38]. The tomo- graphic atom probe holds the tantalizing potential of determining two- or three-dimensional chemical maps on the OS-I.0 nm scale (391. This may ulti- mately enable us to compare MC simulations with experiments on an atom-by-atom basis.

GBs are two-dimensional structures that consist of one or more interface phases, and thus phase tran- sitions are possible. It is only recently, however, that the study of GB phases and their transitions has occurred. There are two basic reasons for this. First, experimentally it is considerably more difficult to access a buried GB than a free surface. And, second, the number and type of processes that can occur at a GB is vast and difficult to quantify due to the larger number of thermodynamic variables. To elaborate, the GB has a local phase rule that includes the five geometric DOFs. For example, for the case of a GB in thermodynamic equilibrium with single-phase grains, this phase rule is given by 1401:

+=c+7--(p; (4)

where JI is the number of independent variables, cp the number of GB phases, C the number of com- ponents in the system and 7 is T, P and the five macroscopic DGFs of the GB. Equation (4) demon- strates that the exploration and construction of GB phase diagrams is a formidable experimental chal- lenge as one needs to explore a C + 6 dimensional phase space.

In spite of the complexity of interface phase dia- grams a number of transitions at GBs have been enumerated theoretically or observed experimentally. Cahn classified GB transitions as either congruent or noncongruent. For the former, the transitions occur with no change in boundary orientation-the inter- face remains planar with its core structure changing. For the latter, a change in boundary orientation is requisite. Few boundary orientations transform con- gruently, thus, an understanding of, for example, the faceting transition that describes a change in GB orientation, becomes important. In particular, this

noncongruent transition results from a change of the inclination of the original interface plane with all other thermodynamic variables fixed [41]. This tran- sition, also denoted roughening, is analogous to the roughening transition observed at free surfaces [42]. GB roughening can be induced by heating a specimen with a faceted interface [43]. Phase transitions pro- ducing new structures can also be induced by segre- gation of solute atoms to an interface. For example, the segregation of Nb to stacking faults in Co pro- duces two coexisting phases in the plane of the fault-a random two-dimensional solid-solution and a two-dimensional ordered phase Co,Nb [44]. Also, the segregation of Au, Sb or S solute atoms to (002) twist boundaries in Fe results in the formation of a new network of dislocations; their cores possibly act as preferential sites for Au segregation [45,46]. GBs thus exhibit many fascinating and intriguing transitions.

The subject of interface phases and their transitions is well developed for solid/vacuum surfaces, and there are many analogies between the phenomena occur- ring at GBs and free surfaces [47,48]. For example, dissolution of a solute [49] at a metal/vacuum inter- face (surface) creates an interfacial region whose chemistry differs from that of the bulk. Pyramidal superstructures of solute, analogous to those that we observe at GBs via MC simulations, are calculated. Although the kinetics of dissolution has been the subject of theoretical work, little accompanying experimental research exists [S&52]. In order to search for correlations between dissolution and GB segregation on an atomic scale it is necessary to employ APFIM to chemically map this interfacial region.

First, in this paper we focus on a methodology to measure experimentally the value of F2. Our ap- proach involves a combination of TEM to measure the five macroscopic DOFs and APFIM to determine F2 directly for the same GBs-see Section 2. An application of this methodology is presented for a series of GBs in a dilute Fe(Si) alloy. It is demon- strated that this approach allows one to directly access GB phase space. Second, an approach is presented for studying GB segregation employing Monte Carlo (MC) simulations-see Section 3. The results of a MC study of solute-atom segregation in the Ni-Pt system are presented for a series of (002) twist boundaries. The emphasis is on extracting a detailed atomic scale picture of the segregation patterns, and how they are affected by the twist angle. In this manner a trajectory in GB phase space is accessed. Both approaches are shown to yield new and unanticipated information about GB segre- gation.

Page 4: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1038 D. N. SEIDMAN et al.

2. ABSOLUTE ATOMIC SCALE MEASUREMENTS OF THE GIBBSIAN INTEBFACUL EXCESS OF

SOLUTE AT GRAIN BOUNDARIES

2.1. Introduction

The methods widely used to study segregation involve exposing a GB surface by in situ fracturing of a polycrystalline or bicrystal specimen. This is limited to metals that fail intergranularly, and, moreover, the atomic arrangement at a GB is necessarily disturbed; also the orientation relationship is lost after fracture. Furthermore, techniques applied to solid/vacuum interfaces, for example, Auger electron spectroscopy, do not yield direct quantitative measures of Tz. Other techniques that indirectly measure Fz involve measur- ing the free energy of a GB as a function of compo- sition and then applying the Gibbs absorption isotherm equation [53,54]. The latter investigations are restricted to high temperatures, and the identity of the segregant must be known a priori. Because of these limitations, our studies are the only ones that include all the state variables required to properly characterize equilibrium GB segregation in specific alloys. Only the APFIM technique directly accesses GB chemistry with single atom detection capability

WI. We employed APFIM to measure directly equi-

librium segregation, as a function of T, at stacking faults [56]. To apply this approach to other internal interfaces requires the use of APFIM in com- bination with TEM-APFIM/TEM. TEM yields the geometric DOFs of the more general type of GBs [57]. First, we identified key factors required to analyze a GB via APFIM/TEM [58]; we thus demonstrated via the W(Re) [59] and Fe(Si) [60] alloy systems that a wide range of GB types can be systematically analyzed. Second, we demonstrated that the atom probe technique can be used to measure directly f,. We also demonstrate that segregation profiles can be measured with atomic scale res- olution.

The basic theory and procedures for studying heterophase or GB segregation are presented. We review our approach to measure Tz as a function of the five macroscopic DOFs using a combination of TEM and atom-probe microscopy and present an application to the study of GB segregation in a dilute Fe(Si) alloy.

2.2. Atom-probe$eld-ion microscope determination of the Gibbsian interfacial excess of solute at grain boundaries

2.2.1. Theory of the measurement. We begin by considering the Gibbsian interfacial excess of element i, Ti, for a heterophase interface as illustrated in

Fig. 1 and then present the special case of a GB. For a heterophase interface Fi is defined as:

N’-= 1 Fi = -.!..-

A =--(N~‘-N;-N~)

= f NVO’[C$’ - cyc - Cf(l - r)], (5)

where NT is the excess number of atoms associated with an interface, and A is the interfacial area over which F, is determined. The quantities Nq and Nf are the number of atoms of element i in phases a and /3, assuming that the two phases exist up to the Gibbs dividing surface, c. The excess is, therefore, defined by comparing the total number of atoms of element i in the actual system containing the interface (NT’), to a comparison system where the two phases making up the interface extend to the position of [. The quantities C; and Cf are the atomic concentrations of element i in the homogeneous regions of phases a and B, that is, the bulk regions of the two phases. The quantity CT’ is the atomic concentration of element i in the total volume of material containing the internal interface, and N”“’ is the total number of all elements in the same volume.

The quantity Ti is defined such that its value depends upon the position of the dividing surface [61]. This dependence is eliminated by defining a so-called relative Gibbsian interfacial excess of com- ponent i with respect to component 1, Tj’):

For a homophase bicrystal, eqn (6) becomes:

where Ca is the atomic concentration of element i in the bulk. For a substitutional binary alloy, under the condition of a dilute solution, F$‘) x r2, and this is the result that we apply to GBs in Fe(Si).

When the APFIM is used to chemically analyze a heterophase interface, a volume of material is probed as in Fig. 1. The APFIM determines the chemical identity of individual atoms, one at a time, in a cylinder of material on an atomic layer-by-layer basis. The spatial and temporal order in which the atoms are detected is preserved. This data is then plotted in the form of an integral profile, as shown schematically in Fig. 2, which is a plot of the cumu- lative number of atoms of element i versus the cumulative number of all of the atoms collected. The

Page 5: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Soluteatom segregation at grain boundaries 1039

Fig. 1. Three-dimensional schematic drawing exhibiting an atom-probe field-ion microscope specimen containing a heterophase internal interface. This heterophase interface separates two different phases (a and b) with the atomic concentrations C; and Cf of element i in the homogeneous regions of phases a and 8, that is, the bulk regions of the two phase phases. The quantity CF’ is the atomic concen- tration of element i in the total volume of material contain- ing the heterophase interface. The unit vector n is normal to the plane of the interface, I is the unit vector parallel to the direction of the atom-probe analysis, and rj is the angle between the vectors n and 1. The maximum depth resolution

of an integral profile is achieved when 4 is zero.

data is plotted in the order in which the ions are collected; thus an integral profile constitutes a spatial and temporal record of the field-evaporated ions. The slope between any two points, therefore, corresponds to the average atomic composition of element i of a particular plane or series of planes in the volume of analyzed material. From the integral profile, we can directly determine the value of NY-“, using eqn (5). The values of NT and Nf are determined by linearly extending the average bulk compositions of the a- phase and the B-phase to the Gibbs dividing surface.

N-

CUMULAm NUMBER OF ALL ATOMS

Fig. 2. A hypothetical integral profile (line ABCD) of a specimen containing an internal interface. All the relevant quantities for determining ri are indicated. See text for a definition of each term. The quantity ri is determined via the atom-probe technique by measuring the chemical identities of all the atoms contained in the indicated volume of

material-see Fig. 3.

The quantity NT’ is the total number of atoms plotted on the ordinate. This graphical construction is simple and does not require any assumptions. The value of A is given by:

AD& A=-,

4cusqJ

where D,, is the diameter of the probe projected onto an FIM surface. The value of A is determined by projecting the probe hole onto a GB plane; this projection is, in general, an ellipse, thus:

r =4cos# i rtnD PC“’ - N: - Nf1

P

4cosqI = rtkD2,, N’“‘[(C~‘) - <c:)t - (cf)(l - t>l.

The atomic concentrations are expressed as averages, because the APFIM measures these averages. The quantity q is the detection efficiency of the chevron detector of the APFIM. The value of q equals the fractional open area of a chevron detector and is typically 0.55. The value of 0 is given by:

(fB = co-‘(n~ I), (10)

where II is the unit normal to the interface plane; I is the unit vector parallel to the direction of APFIM analysis, and is determined by indexing the center of an FIM image of the projection of a probe hole. The image is indexed by determining the rotation matrix from a TEM analysis (621. The value of D,, is then determined directly from information in an FIM image. For the experiments reported here, the value of cos 4 is unity, as P is parallel to I and therefore the projection of the probe hole is a circle.

For a homophase bicrystal the bulk compositions on each side of a GB are identical, so that the location of a dividing surface does not influence the value of N;“-. For the Fe(Si) alloy, experimental measure- ments show, however, that the bulk compositions do vary slightly due to the statistical nature of a compo- sitional analysis. To account for these. variations a Gibbs dividing surface is placed at the center of the region that constitutes a GB. This placement is equivalent to using the average of the bulk compo- sitions of both grains for the value of C;” in eqn (6).

The schematic diagram in Fig. 3(a) shows an atomic scale picture of the analysis for a GB in an Fe(Si) alloy; this figure corresponds to the situation where II is parallel to 1. To extract Tsi, the excess number of Si atoms, N,i, in the volume and area,

Page 6: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1040 D. N. SEIDMAN er al.

(a) @) cIlmmeleteclnm PmlX n$W=“Y /hole La

l Soluteatam-Si

=tvzad 0 Solvent atoms -Fe

Fig. 3.(a) Ideal geometry for an atom-probe analysis of a GB. The projection of the probe hole onto the surface of the tip of a specimen determines the area over which TS is measured. (b) Volume of material analyzed. For the experiments presented the direction of analysis (l) is parallel to the unit GB plane normal

(n).

A, over which it is determined, is measured; fsi is given by:

4cosl$ = rlxD2, N’“‘[(C$) - (C:‘)<

P

- (Gi22U - 019 (11)

where the quantitites Ng’ and NE* are the number of Si atoms in grains 1 and 2; the two grains exist up to a Gibbs dividing surface, <. The quantities (Cp' )

and (C$‘) are the average atomic concentrations of Si in the homogeneous bulk regions of grains 1 and 2; (C$“) is the average atomic concentration of Si in the total volume of GB material. All the quantities in eqn (11) are indicated on Fig. 3(b).

double-tilt stage for an Hitachi H-700H 200 kV trans- mission electron microscope; this stage is vibra- tionless at a magnification of 310,000 x . It has a tilting range of f 30” for the x-tilt and + 27” for the v-tilt; this range is sufhcient to analyze crystallo- graphically an internal interface. This double-tilt stage and transmission electron microscope are also used to examine atom-probe specimens during the course of a backpolishing treatment to place selec- tively an internal interface in an FIM tip.

We have employed this approach extensively to study segregation at GBs in binary metal alloys and presently in Ni-based intermetallics. If, however, the interfacial area per unit volume is sufficiently large (> 10gm2m-3) it is possible to perform a random area analysis to locate an internal interface; this obviates the need to use TEM to place an interface in the field of view.

2.2.2. Locating an internal interface for atom -probe 2.2.3. Measurement of the Gibbsian interfacial ex-

spectroscopy. The general approach to observe a cess of Si at GBs in an Fe(Si) alloy. First, the study specific internal interface is to prepare an FIM spec- of equilibrium solute-atom segregation in an Fe(Si) imen from a bulk specimen that was processed to alloy requires that the starting materials must have produce a desired microstructure. The FIM specimen utmost purity to minimize the influence of impurity- is then prepared by electropolishing, electroetching, atom segregation of the interstitial elements carbon, ion-beam milling, or a combination of those pro- phosphorous, sulfur, nitrogen or oxygen. Thus the cesses. One specific approach we employ involves starting materials were Fe of 99.9945 at.% purity and electrolytic backpolishing to an internal interface, Si of 99.9999 at.% purity. The Fe-3 at.% Si alloy was employing a versatile system for systematically arc melted in an atmosphere of argon of high purity. preparaing atom-probe specimens. This systematic The ingot was swaged and cold drawn to 185pm approach incorporates a.c. electroetching or d.c. diameter wire. Next the specimens were encapsulated electropolishing in automated or manual modes [63]. in a quartz tube backfilled with Ar of high purity. The a.c. waveforms available are sine or triangular or Then they were annealed at 823 K for 14.5 or 72.5 hand square in either the one shot or continuous wave quenched into a brine solution at 273 K. Two anneal- modes. The d.c. electropolishing mode produces ing times were used to ascertain that the equilibrium pulses with widths in the range from OSms to concentration of Si was achieved at the GBs. These 500 s. The power supply provides 0 to +48 Vd.c. at annealing times produce root-mean-square diffusion up to 1 A. distances of 0.055 and 0.123 pm, respectively.

In our laboratory an atom-probe specimen is examined by TEM utilizing a radically modified

Figure 4(a) exhibits a transmission electron micro- graph of a GB near the tip of an Fe 3 at.% Si wire

Page 7: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Solute-atom segregation at grain boundaries 1041

0 a

Fig. 4.(a) A transmission electron micrograph of an Fe-3 at.% Si specimen after GB analysis and backpolis~ng. (b) A field-ion microscope image of the same specimen as in (a). The tip has been field evaporated so that the image of the GB perimeter is circular. The probe hole is aligned so that its perimeter is concentric with that of the GB image: C$ = 0” and D h = II.5 nm. This Fe-3 at.% Si specimen had been

annealed at 823 fc for 14.5 h.

that was backpolished to produce a tip with the the abscissa represents the c~ulative number of

requisite geometry to locate a GB and analyze it via Fe plus Si atoms collected from a cylinder of

atom-probe microscopy. Figure 4(b) shows an material 11.5 nm in diameter. A pulse fraction (f)

FIM micrograph of the same tip. No signs of precipi- of 0.15-j” is the ratio of VP to V,,C, -

tation are observed in the bulk or at GBs by either and a specimen temperature of 35 K were

transmission electron or atom-probe microscopy. employed (see Krakauer and Seidman [64]). The

Figure 5 is the corresponding integral profile; value of rsi is 0.378 If: 0.12 x 1Ol4 atoms cm-*.

1.0 0.8 0.6 0.4 0.2 0 0.2 0.4 / ‘

Fc-3at.%Si tn 360 L-_S=O.665

B T, =823 K I t,=14.5h Tsg =0.378 f0.12xlO” akoma cm** j

Grain -4 Boundary

4.2f0.21 at.% Si

* 0 wo lSo0 27m 3600 4500 5400 6300 7200 8100 !a00

CUMULAm NUMEER OF Fe PLUS Si ATOMS

Fig. 5. Integral profile of the GB analyzed in Fig. 4. The value of NV” is 8915 atoms. The pulse fraction (V,JV,,C,) employed was 0.15, the temperature was 35 K, and the pulse frequency was 60 Hz.

Page 8: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1042 D. N. SEIDMAN et al.

The five macroscopic geometric degrees of freedom are: c = [0;85,0.52,0.06], 6 = 25.28”, ni = --- [O&l, 0.68,0.73], and n, = [0.21,0.36,0.91]. Figure 6 is a layer profile calculated from the data displayed in Fig 5. Each silicon concentration data point plotted on the ordinate is based on a total of 300 atoms/layer; note that the data points are 0.05 nm apart, and that there is a slight depletion of silicon on the left-hand side of this GB. This type of profile is the experimental version of the segre- gation profiles calculated via MC simulations, see Section 3.3.

The depth scale on the upper abscissa of Fig. 5 is determined from a knowledge of i), and the atomic density of Fe (pi+ = 84.879 atoms nm-‘); note well that each tick mark corresponds to 0.2nm. The ultimate depth resolution of an atom-probe compo- sitional analysis is equal to the mean interatomic planar spacing (d,,,) along II, and is approximately equal to 0.02 nm. Because the probe hole covers approximately 3.5 atomic planes, the actual depth resolution is xO.07nm. Since the vector n is not along an iden~able low-index pole in the FIM image, the Miller indices can only be assumed for this atom-probe analysis. An example of an easily identifi- able pole is 110 for which the depth resolution is 0.02 nm. The error bar for Tsi is determined by propagating the errors of each quantity in eqn (6). The error in NFexcesS is determined by assuming that the measured atomic concentrations are binomially dis- tributed 1651. The error in I)ph is ernpi~~lly deter- mined to be f7%. The value of rS corresponds to a fractional monolayer coverage of 8 = 0.22 mono- layers by assuming a saturation value for rsi of p,,d,,, ,

which equals 1.70 x lOi atomscmW2. If we assume that the solute-atom segregation behavior obeys a Langmuir-McLean type adsorption isotherm, then the free energy of segregation for this GB is - 15.1 kJ mall’.

2.2.4. The grain boundary phase space of an Fe(S) alloy. Thirteen GBs were analyzed as described in the previous section. As rsi is a function of c, 8, and n, at fixed T, P, and bulk composition, the data spans a six-dimensional space. In Fig. 7, five DOFs are folded into two axes to display rsi in a three-dimen- sional plot. Four of these DOFs are folded into the abscissa cos( < e, II), while the fifth DOF, the disori- entation angle, @,,,, is displayed along the second axis. This corresponds to the degree of twist or tilt of a GB. Figure 7 shows that rsi rapidly increases from 0 at &, = 0” to 1 .O x 1Or3 atoms cmm2 at the transition from low- to high-angle. Beyond this transition, Tsi increases to 2.5 x IO” atomscm--’ (ediS = 68”). (For sin((Idi,/2) = 0 the value of Tsi is zero.) Also, the GBs with c. n = 1 (pure twist) have higher values of Tsi than do CBS with c. n = 0 (pure tilt). This demon- strates that twist GBs exhibit a higher level of segre- gation than do tilt GBs in this alloy. This is the ftrst experimental evidence that twist and tilt GBs exhibit different absorptive capacities for solute atoms. We have recently obtained some evidence for this same fact employing Monte Carlo simulations on E = 5 twist and tilt boundaries for a dilute Pt(Au) alloy [66].

1. A direct, quantitative and absolute method- ology has been developed to determine the Gibbsian interlacial excess of solute, TZ, at preselected GBs.

DEPTH (nm)

Fe-3at.%Si

T, =823 K

t,=14.5 h

01 .I. t 0 10 20 30

NUMBER OF ATOMIC LAYERS (300 ATOMS/LAYER)

Fig. 6. This is a layer profile calculated from the integral profile displayed in Fig. 5. A dotted line is placed at the position of the peak concentration at the GB. The schematic diagram in the upper right hand corner

indicates the procedure used for the analysis.

Page 9: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Solute-atom segregation at grain boundaries 1043

Fig. 7. A three-dimensional bar chart of the Gibbsian interfacial excess of silicon, Tsi, vs sin(0,,/2) and cos( ~c, n). The height of each vertical bar is equal to the value of Tsl. The bars arc shaded differently

to make them easier to distinguish from one another.

2. The approach involves the combined use of TEM and atom-probe microscopy-APFIM/TEM. TEM is used to first determine the five macroscopic degrees of freedom, DOFs, of a GB in an atom probe specimen, employing a specially developed double-tilt stage that holds 1 cm long field-ion microscope wire specimens. The same specimen is then electrolytically backpolished to the GB, and transferred to an atom probe.

3. We have applied this methodology to GBs in

single-phase alloys of the W(Re) and Fe(Si) systems. 4. Results are presented for a high-purity Fe-

3 at.% Si alloy that was recrystallized and annealed at 823 K for 14.5 or 72.5 h; these annealing times correspond to root-mean-squared diffusion distances of 0.055 and 0.123 pm respectively.

5. For this Fe(Si) alloy Tsi was measured for GBs with different structures. A three-dimensional plot of Tsi vs cos( LC, n) and sin(B,J2) is used to present the results; in this manner all five macroscopic DOFs are folded into two axes-see Fig. 7. This figure consti- tutes the phase space of GBs for this alloy.

6. Figure 7 shows that for large values of f& that

Tsi saturates for all the GBs studied. Although the saturation value depends on the character of the GB. In particular, for this alloy the saturation value is larger for pure twist boundaries, cos( LC, n) = 0, than it is for pure tilt boundaries cos( LC, n) = 1. This is the first experimental measurement that demonstrates that there is a difference in the absorptive capacities of tilt and twist boundaries.

7. Atomic scale segregation profiles normal to the GB plane are measured for specific GBs, see Fig. 6.

3. MONTE CARLO SIMULATIONS

OF SOLUTE-ATOM SEGREGATION AT (002)

TWIST BOUNDARIES IN DILUTE FCC ALLOYS

3.1. Introduction

In spite of numerous experimental, theoretical and simulation studies our basic knowledge about GB segregation is still rather fragmentary. This is a result of the enormous complexity of the 6 + C dimensional phase space in which this phenomenon occurs. Be- sides the conventional state variables-temperature, pressure and composition-the five macroscopic DOFs of a GB are postulated to be thermodynamic state variables. The three microscopic DOFs, that is, the rigid body translation vector between two grains, are assumed to be fully relaxed to minimize the Gibbs free energy of a bicrystal. We have investigated solute-atom segregation in detail at (002) symmetri- cal twist boundaries in dilute Au-Pt and Ni-Pt alloy. And in this section we present an abbreviated account that outlines the important features of segregation in the Ni-Pt system, focusing on low-angle (002) sym- metrical twist boundaries [67]. The Ni-Pt system is of particular interest for several reasons:

1. It exhibits a clear tendency to ordering with three ordered phases (Ni,Pt, NiPt and NiPt,) in the bulk phase diagram [68]; this is in sharp contrast to the Au-Pt system which exhibits a miscibility

gap. 2. There is a considerable size misfit between the

two elements in contradistinction to the elements Au and Pt; the room temperature values of the lattice parameters are 0.352 and 0.392nm for Ni and Pt, respectively, while for Au it is 0.408 nm.

Page 10: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1044 D. N. SEIDMAN et al.

3. Oscillatory solute-atom segregation profiles are observed in Ni-Pt alloys near different free surfaces [69], at a mixed E- = 5 twist-tilt boundary [70], and at an (002) twist boundary.

All the simulations we~~rfo~edat a temperature, T= 850 K, that is high enough to be reproduced experimentally (in terms of diffusion times) but low enough to observe substantial solute-atom segrega- tion effects. The solute-atom concentrations on both sides of the phase diagram,were chosen to lie in single- phase solid-solution regions. The solut*atom con- centrations are high enough to give adequate stat- istics but the alloys can still be considered to be dilute.

The main questions of concern are:

* What is the structural dependence of solute- atom segregation? that is, the dependence of the Gibbsian interfacial excess of solute on the twist

angle of the boundaries. l What is the detailed spatial arrangement of

solute atoms at a grain boundary as a function of twist angle for low-angle boundaries?

e How is the solute-atom segregation behavior related to the the~odynamic properties of the alloy system?

3.2. Computational procedures

The computational methodology employed is MC simulation, which is well suited for atomistic simu- lations of equilibrium phenomena in solids [71,72]. We utilize the fundamental algorithm of Metropolis et al. [73] as it provides a reasonably good conver- gence to equilib~um and efficient sampling of the ~uilibrium wn~~rationa1 space. The choice of the statistical ensemble was dictated by the necessity to simulate a GB in a single-phase solid solution that is in equilibrium with a quasi-infinite reservoir of solute atoms in a bulk reference crystal. The most suitable approach for this situation is the so-called transmuta- tional ensemble. The parameters held constant are temperature, the total number of atoms and the difference in the so-called excess chemical potentials [74]. Atoms are, however, allowed to change their chemical identities, so that the atomic fraction of each component is variable. In addition, the volume of the simulation cell is allowed to relax, thereby yielding a constant pressure.

At each MC step an atom is chosen at random and a decision is made whether or not a certain attempt is accepted-the latter is based on the Metropolis algorithm. The following sequence of attempts is performed:

1. The chemical identity of an atom is changed; this step is for m~ifying the local composition of an

alloy, and it is basic to the phenomenon of solute- atom segregation.

2. The displa~ment of an atom from its position by a random vector with a magnitude smaller than 0.01 nm; this step allows for relaxations on an atomic scale. Note well that steps 1 and 2 are made simul- taneously.

3. After, on the average, all the atoms in a bicrystal have undergone the above two attempts an overall volume relaxation is implemented that involves a change in one or more periodic lengths, and a co~esponding resealing of wordinates of all the atoms in the bicrystal by a small random quantity. This procedure changes the volume of the compu- tational cell to maintain the bicrystal at constant pressure.

4. A relative translation of the two grains in the plane of the interface by a random vector is at- tempted with the same time periodicity as step 3. It allows complete relaxation of all microscopic degrees of freedom of the boundary. No significant displace- ments from the coincident site lattice (CSL) structure were, however, observed for the low-angle bound- aries.

The total value of the internal energy E,, is calcu-

lated using the embedded atom method (EAM) po- tentials that have been developed for six face-centered cubic elements (Ag, Au, Cu, Ni, Pd and Pt). The EAM potentials are empirical many-body continuous potentials that have been used extensively for study- ing different physical problems in materials science [75-781.

Prior to simulations for GBs, similar simulations are run for a perfect single crystal (4000 atoms) to determine the value of Ap---the difference in the excess chemical potentials--corresponding to the req- uisite value of solute-atom concentration. The equi- librium lattice constants for given concentrations are also obtained from those runs (lo4 MC steps per atom).

The computational cell represents a bicrystal with thr~-dimensional periodic boundary conditions. The two grains are rotated about an [OOl] direction to obtain a given twist angle; hence, only one of the five macroscopic geometric DOF is varied. Physically, there are two crystallographically identical GBs sep- arating the two grains. This arrangement is valuable because it eliminates the free surface, and improves the statistics by having two GBs in a given bicrystal. Also a sufficiently large separation is necessary be- tween the two CBS to exclude the interaction of the stress fields of the two GBs in the middle of the bicrystal, where it is necessary to maintain the prop- erties of the bulk single crystal. We find empiricaliy

Page 11: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

that commencing with 16 (002) planes between the two interfaces this condition is met for the lowest- angle boundary-that is, the one having the longest range elastic stress fields-we investigated. Normally in our MC simulations we use a computational cell that is approximately (4-5) nm x (4-5) mn in the plane of the interfaethis corresponds to between 200-300 atoms in one atomic plane-and is about 6.3 nm in the direction normal to the (002) GB plane. Thus, the overall number of atoms in the system varies between 6400-9600. The first 2.5 x lo3 MC steps per atom of each run were used for equili- bration. The averaging was performed over the next 3 x lo3 or 104 MC steps per atom, depending on the boundary. The time step for averaging was of the order of 10 MC steps per atom to avoid temporal correlation effects.

Pt-3 at.% Ni w-3 at.% Pt

0.14 - e=5”

0.08.

0.08

Another impo~ant effect is the surface tension of a GB that acts in the plane of an interface to reduce its cross-sectional area. This effect causes stresses that cannot relax locally-due to the periodicity in the plane of the interface--so that the entire bicrystal has to accommodate these stresses. In our case the stresses are accomodated by means of a lattice expan- sion normal to the GB plane. The magnitude of the effect is approximately proportional to the fraction of the total system volume belonging to the GB phase; that is, essentially the inverse of the distance between the two GBs in our bicrystals. The typical “thickness” of a GB is approximately one to two atomic planes. Thus in a normal polycrystalline specimen this effect most likely does not exist; it may, however, exist in nanograined material. In a MC simulation this effect may be quite large if a bicrystal is too small. A

2 8 16 24 32 8 16 24 32

(002) planes a

Pt.3 at.% Ni Ni-3 at.% Pt

0.14 I 9=28.10 t 8

0 = 28.1” t

3

s 0.14

2 0 J 0.08

Y &

I T = 850 K

I”“,, , .,- 2 Ni-3 at% Pt : i 0.0% 0.14

Fig. 8. A plot of the Gibbsian interfacial excess of solute (I’,) vs sin(e/2) for Pt-3 at.% Ni (open squares) and Ni- 3at.% Pt (solid circles> at 85OK. The quantity sin(@/2) is directly proportional to the primary grain boundary dislo- cation density via the Read-Shockiey equation. The error

bars indicate plus or minus one standard deviation.

Solute-atom segregation at grain boundaries 1045

1 8 16 24 32 8 16 24 32

(002) planes

b

Fig. 9. Solut*atom segregation profiles normal to the geometric (002) interface plane for the nine boundaries studied at 850K. The left-hand column corresponds to Pt-3 at.% Ni and the ~ght-hand one to Ni-3 at.% Pt. Each pfot contains two c~s~lograp~~ly identical periodic interfaces-their planes lying between planes M-17 and 32-l. (a) For the twist angles (6) So, 10.4”, 16.3” and 22.6”. (b) For B = 28.1”, 33.9”, 36.9” 41.1” and 43.6”. The error bars correspond to plus or minus one standard deviation.

Page 12: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1046 D. N. SEIDMAN et al.

straightforward but computationally inefficient sol- (r: = 17) 33.9” (;r, =289), 36.9” (z: = 5), 41.1” ution is to increase the separation between the GB (Z = 73) and 43.6” (Z = 29). The choice of CSL planes. We employ an alternative solution, where boundaries is not restrictive, because the boundaries the total area of the interface is held constant at cover the entire range of possible r&orientations. a given ~~lib~~ value, the latter is dete~in~ The atomic structures of the (002) grain boundaries by the equilibrium value of the lattice parameter have been extensively studied experimentally by X- found from simulation runs for a single crystal. Thus, my diffraction techniques, geometrically by the pack- only the periodic length normal to the interface is ing of hard spheres [79], anisotropic elasticity theory allowed to relax. We have demonstrated that either [80], lattice statics [81,82], molecular dynamics [83] employing much larger bicrystals-40 (002) planes and free-energy minimization [84] techniques. between the GBs-or fixing the cross-sectional di- Relaxed low-angle (002) twist boundaries in cubic mensions of an interface yields the same quantitative metals consist of a square grid of localized orthog- results. onal screw dislocations. The resulting twist boundary

The bicrystal employed is one where the [OOl] is in a state of pure shear with no long range stress direction is perpendicular to the plane of the (002) fields [SS]. This description appears to be an excellent interface plane; the geometric interface lies between one for twist boundaries in metals. The screw dislo- two (002) planes and it contains no atoms. All cations are the so-called primary grain boundary possible (002) twist boundaries lie in the angular dislocations (PGBDs) whose Burgers vectors (h) are range 0 = 0” to 45” about the [OOl] direction; the (a/2)<1 lO)-type (a is the lattice parameter) in the point group for this interface is 4mm. Of the five face-centered cubic lattice. The spacing of the grid is macroscopic geometric degrees of freedom only 6 is given by the Read-Shockley relation: varied. The angular values studied correspond to the following CSL boundaries: 5” (X = 265) 10.4” (C= 61), 16.3” (X= 25), 22.6” (z: = 13), 28.1”

5nm 4 + ~~001

(12)

Fig. 10. Patterns of Ni atom segregation at the four successive (002) planes nearest the geometric interface for a @ = 5” twist funds in a R-3 at.% Ni alloy. Atomic sites with au average Ni ~n~ntration that is one or more standard deviations above the bulk Ni concentration are indicated by solid black circles

and the remaining sites are gray circles.

Page 13: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Solute-atom segregation I at grain boundaries

The deviation from an exact CSL orientation pro- duces a grid of secondary grain boundary dislo- cations (SGBDs) with the spacing given by the same Read-Shockley relation, with 0 replaced by IAel; where A0 is the deviation angle measured from the exact CSL orientation [86]. The intersections of the PGBD lines form patches of the Z = 5 structure. The localization of the misfit falls off with the distance from the interface, and the dislocation cores become broader and less pronounced. At sufficiently large angles (0 3 22.6”) the areas of the regions of “good” atomic fit become comparable to those of the inter- section regions, and more complicated high-angle twist boundaries appear that can be described by a structural unit mode1 [87,88]. The atomic displace- ments within the square cells of the PGBDs are rotations around the elements of Bollmann’s O- lattice.

For (002) twist boundaries only symmetrical struc- tures exist. A displacement, however, of the upper grain with respect to the lower grain-parallel to the plane of the interface-may lead to a change of the symmetry of the boundary. For twist boundaries

1047

crystallographically nonequivalent displacements of the two grains are limited to the cell bounded by the two shortest nonparallel displacement shift complete (DSC) vectors, b, and bz. in the boundary plane. Bristowe and Cracker have shown by means of molecular statics at 0 K, for a Z = 5 twist boundary in pure Ni or Cu that the only displacements leading to stable structures are (1/2)b, or (1/2)b, and (1/2)[b, + b,]; these structures are denoted type I and type II. Recently, a free-energy minimization study, in the local-harmonic approximation [89], revealed a phase transition from the CSL to type I structure in a Z = 5 high-angle twist boundary in gold at about 315 K. In our previous papers [90] Monte Carlo simulations were performed for the type I and type II as well as the CSL Z = 5 high-angle boundaries in a Pt-1 at.% Au alloy at 850 K. The type I structure reverted to the CSL structure and the type II struc- ture remained metastable. There were, however, no significant differences between the solute-atom segre- gation profiles of the type II and CSL structures. We have recently discovered a phase transition in the ;I: = 5/(002) twist boundary in the Pt(Ni) system. This

3nm 4 * WOI

Fig. 11. The pattern of Ni atom segregation at the three successive (002) planes nearest to a 0 = 10.4” twist boundary in a Pt-3 at.% Ni alloy. Atomic sites with an average Ni concentration that is one or more standard deviations greater than the bulk concentration are indicated by solid black circles and the

remaining sites are gray.

Page 14: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1048 D. N. SEIDMAN et al.

transition occurs when the bulk concentration of Ni exceeds x6 at.%; it involves a change from the CSL structure to the type II structure [91]. A drastic change in the distribution of solute concen- trations at different GB sites accompanies this tran- sition.

3.3. Solute-atom segregation behavior

As discussed in Section 2.2.1 r2 is the most general measure of solute-atom segregation-it applies to any solid/solid interface. For analyzing the results of the simulations on GBs we define r, in terms of dimensionless units of solute monolayers by:

r, = 2 (Ci - Co), (13) i=l

where npl is the number of (002) planes in the bicrystal and ci and c,, are average solute-atom concentrations, in units of atomic fraction, in plane i and in the bulk of the bicrystal, respectively.

=;

E

1 1 I-

2

The dependence of r, on sin(0/2) is exhibited in Fig. 8 for the Ni-3 at.% Pt and Pt-3 at.% Ni alloys. The dislocation density is simply the reciprocal of d

in eqn (12), hence the quantity sin(0/2) is pro- portional to the PGBD dislocation density. For both alloys 6/2 increases with increasing dislocation den- sity, that is, with increasing 0. The value of Tz increases monotonically for both alloys, and then saturates at about half of the angular interval (6’ > 22.6”). This angle corresponds to the dislocation density at which the dislocation cores start to overlap. The magnitude of the effect is about a factor of two larger for the Ni-rich alloy.

The solute-atom segregation profiles normal to the (002) GB plane yield the next most detailed level of information. These profiles for nine GBs are exhib- ited in Figs 9(a) and 9(b) in order of increasing 0. The left-hand column is for the Pt-3 at.% Ni alloy, and the tight-hand column is for the Ni-3 at.% Pt alloy. The two geometric interfaces are located between the 16th and 17th (002) planes and the 1 st and 32nd (002)

Pt-3 at.% Ni

5.6 nm 4 * WI

a

4 5.6 nm

b

Fig. 12.(a) A projection of atomic sites in a Pt.3 at.% Ni alloy-for a 6 = 5” twist boundary-normal to the (002) interface plane, onto a {llO}-type plane, and to one of the two orthogonal sets of screw dislocations along the (1 IO)-type directions. The distance between the two screw dislocation lines is 11.5]b]. The five first (002) planes on both sides of the geometric interface, denoted I-I, are shown. The details around the second set of orthogonal dislocations have been removed for the sake of clarity, so that the regions adjoining the cores of the two parallel screw dislocations--normal to the plane of the interface--show up. The sites enhanced in solute are depicted by large spheres, while all the other ones by small points. The enhanced sites are arranged in hourglass-like patterns above and below the dislocation lines. The solute-atom concentrations at individual enhanced sites decrease with increasing distance from the interface. The rotation angle between the upper and lower grains is seen from the projections of atomic rows. (b) The same picture rotated by 6” around an [OOI] axis normal to the interface

to exhibit the three-dimensional nature of the pattern.

Page 15: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Solute-atom se~egation at grain boundaries 1049

planes; the presence of two GBs is due to the periodic boundary conditions. The 1st to 16th (002) planes belong to one gram of the bicrystal, while the 17th to 32nd (002) planes belong to the other grain. So- lute-atom segregation is expected to be symmetrical with respect to the (002) interface planes because the GB is a symmetrical twist boundary. The concen- tration values plotted on the ordinate scale are the average solute-atom con~entra~ons integrated over the area of each (002) plane. The error bars represent pfus or minus one standard deviation, and are of different magnitudes for different GBs because of different durations of the simulation runs. The planes in the region between the two interfaces constitute the bulk of a bicrystal, and they are unaffected by solute--atom segregation; this region has a constant solute-atom concentration that corresponds to the butk concentration of each alloy. The solute concen- tration is highest near the interface, with the excep- tion of the B = 5” boundary in the Ft-3 at.% Ni alloy; this boundary exhibits a Ni con~ntration in the first (002) plane adjacent to the interface that is slightly lower than in the second (002) plane. The solute-

atom concentrations in the (002) planes adjacent to the interface increase steadily until saturation occurs at B ~22.6”. Simultaneously the width of the segre- gation profiles decreases with increasing twist angle; the profiles are fairly sharp for the Pt-3 at.% Ni alloy, and even sharper for the Ni-3 at.% Pt alloy.

Now we examine visually the spatial distribution of atomic sites enhanced in sofute atoms near the inter- face for low-angle boundaries. Figure IO displays the first four @@2) planes adjacent to a @ = 5” twist boundary in a Pt-3 at.% Ni alloy. The atomic sites with a solute-atom concentration greater than one standard deviation above the equilibrium bulk value are black while the remaining sites are gray. The sites enhanced in Ni atoms are situated predominantly in rows along the dislocation cores-the (l/2)( 1 lO>- type directions--and the region of enhanced solute concentration widens with increasing distance from the (002) interface. Note carefully, however, that the average solute-atom con~nt~tiuns at the enhanced sites decrease concomitantly with increasing distance from the (002) interface. The sites in the middle of the dislocation grid are essentially unaffected by

Fig. 13. The pattern of F’t soiute-atom segregation at the four successive (002) pfanes nearest to the B = 5” interface in a Ni-3 at.% Pt aiioy. Atomic sites with an average Ft concentration that is greater than the bulk value by more than one standard deviation are indicated by solid black circles and the remaining

sites are gray circles.

Page 16: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1050 D. N. SEIDMAN et al.

the process of solute-atom segregation. A similar plane has a smaller number of enhanced sites than the physical picture is found for the 6 = 10.4” boundary second one, while for the third (002) plane the (Fig. 11). increase in the number of enhanced sites is compen-

To better visualize the spatial distribution of en- sated for by a decrease in the average concentrations hanced sites a projection onto a (1 IO)-type plane at those sites-the average solut*atom concen- normal to the plane of the interface is shown in tration decreases with increasing normal distance Fig. 12(a). The regions with the dislocations parallel from the (002) interface. Figure 12(b) is the same to the plane of the projection have been deleted, so image as in Fig. 12(a) but it is rotated by 6” around that only the two dislocation cores normal to the the [OOl] direction normal to the (002) interface plane, plane of the projection remain. Five (002) atomic denoted by I-I, to display the three-dimensional planes on both sides of the geometric (002) interface, character of the solute-atom distribution. denoted by I-I, are displayed with the sites enhanced For the Ni-3 at.% Pt alloy two-dimensional plots

in solute denoted by large spheres, and those un- of the solute-atom distribution in the first three (002)

affected by solute-atom segregation by points. Due to planes parallel to the geometric interface are dis- the small (0 = 5’) misorientation between the upper played in Figs 13 and 14 for the 0 = 5” and 0 = 10.4”. and the lower grains, the points representing rows of For this alloy the solute atoms tend to segregate to atoms are slightly shifted in the (002) planes normal regions that are complementary to those in the to the interface. The cross-section of the atomic sites Pt-3 at.% Ni alloy. Rows of atoms along the dislo- enhanced in solute atoms has an hourglass-like shape cation cores are virtually unaffected by solute-atom for both dislocations; the same description also ap- segregation, while the areas inside the dislocation grid

plies to the two dislocations not displayed. This figure are enhanced in solute atoms. Away from the (002) provides an explanation for the oscillatory segre- interface plane the rims of the unenhanced regions

gation profile in this boundary (Fig. 9). The first (002) broaden and the enhanced sites move toward the

Fig. 14. The pattern of Pt soluteatom segregation at the three successive (002) planes nearest to the 0 = 10.4” interface in a Ni-3 at.% Pt alloy. Atomic sites with an average Pt concentration that is greater than the bulk value by more than one standard deviation are indicated by solid black circles and the

remaining sites are gray circles.

Page 17: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Ni-3 at% Pt

4 5.7nm

b

-1

.[llO]

-I

El 101

Fig. IS.(a) A projection of atomic sites ia a Pt-3 at% Ni alloy-for a 0 = 5” twist boundary-normal to the (002) interface and to one of the two orthogonal sets of screw dislocations along the (llO>-type directions; this is similar to the projection in Fig. 12(a). The distance between the two screw dislocation lines is 11.5lb& The first six (002) planes on both sides of the interface, denoted I-I, ate displayed. The sites enhanced in solute are depicted by large red spheres and all the others by small points. The black cotor corresponds to solute-atom ~ncen~ations less than the bulk concentration plus one standard deviation, that is, less than 4at.% solute. The concentration of solute at enhanced sites (with concentrations of solute greater than 4at.%) is represented by the density of color. The darkest red corresponds to concentrations of solute in excess of lOat.%. The enhanced sites are arranged in bip~da~ patterns above and below the screw dislocation lines. The solute concentration at an individual enhanced site decreases with increasing distance from the interface. The rotation angle between the upper and lower grains is seen from the projections of atomic rows. (b) The same picture rotated by 9’ around an [OW] axis normal to the interface to exhibit the three-dimensional character of the pattern.

1051

Page 18: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Solute-atom segregation at grain boundaries 1053

center of the square grid of dislocations. When projected onto a { 1 lO}-type plane, in the same man- ner as in Fig. 12(a), the enhanced regions are found to form bipyramidal regions centered on the same square grid of dislocations-see Fig. 15(a). For better resolution in Fig. 15(a) and (b) different levels of solute concentrations are represented by a color scale. The first six (002) planes on both sides of the inter- face, denoted I-I, are displayed. The sites enhanced in solute are depicted by large red spheres and all the others by small points. The black color corresponds to solute-atom concentrations less than the bulk concentration plus one standard deviation, that is, less than 4 at.% solute. The concentration of solute at enhanced sites (with concentrations of solute greater than 4 at.%) is represented by the density of color. The darkest red corresponds to concentrations of solute in excess of 10 at.%. The rotation angle between the upper and lower grains is seen from the projections of atomic rows. Figure 15(b) is the same picture as in Fig. IS(a) except that it is rotated by 9” around the [OOl] direction normal to the (002) inter- face plane to exhibit the three-dimensional character of the solute-atom distribution.

3.4. Discussion and summary

A systematic investigation has been performed of the phenomenon of solute-atom segregation at (002) symmetrical twist boundaries in dilute single-phase solid-solution alloys on both sides of Ni-Pt phase diagram at 850 K. The following is a discussion and summary of the results presented in Section 3.3.

3.4.1. Chemistry of the alloy. The interfacial re- gions are enhanced in solute atoms on both sides of the Ni-Pt phase diagram. From the point-of-view of classical thermodynamics of the Gibbsian interfacial excess of solute is defined by eqn (2). The sign of r2 can be either positive or negative, depending on the fundamental thermodynamic properties of the solid- solution. A major challenge is to be able to infer at least the sign of TZ from the properties of the bulk phase diagram. One of the simplest suggestions is that the pure element with the lower surface free energy is enhanced at a GB. This implies that the sign of r2 always changes on opposite sides of a phase diagram. This has been found to be true for simulations of solute-atom segregation at symmetrical twist (002) boundaries in the Au-Pt [92], Cu-Ni [93,94] and Au-Pd [95] systems. For all three alloy systems Au and Cu are enhanced on both sides of the phase diagrams, respectively. For all three alloy systems the element with the lower melting point-that is, the element with the lower surface tension segregates to GBs. The present results introduce, however, a differ- ent trend, with the sign of r2 being positive for the

solute species on both sides of the Ni-Pt phase diagram. Other systems with similar behavior are Au-Cu [96] and Cu-Pt [97]; these systems also have large values of the size misfit parameter.

We now attempt to interpret our observations for the Ni-Pt and Au-Pt systems. The real situation may, of course, be much more complicated than the simple model we are presenting. Within the framework of elasticity theory the interaction of a solute atom with a twist boundary [98] arises from two effects. The size misfit effect is a pA V type of interaction, and it arises from local expansions at screw dislocation cores and the resulting net compression is derived from nonlin- ear elasticity theory. The inhomogeneity effect is due to a difference of the elastic constants of the solute and solvent atoms and it comes from linear elasticity theory. Both effects are second order in the distance from a dislocation line. The size misfit effect par- ameter, ca, is proportional to the fractional change of the lattice constant with concentration of the solute species and it is given by ca = a-‘(da/de). The inhomogeneity effect parameter, Ed, is given by L, = p -‘(dp/dc), where p is a shear modulus. Due to the size misfit effect oversized atoms should be re- pelled from twist boundaries, while undersized ones should be attracted to them. Similarly, due to the inhomogeneity effect “harder” atoms should be re- pelled from GBs, and “softer” ones should be “at- tracted”. In the Ni-Pt system, where the size misfit parameter is fairly large and the inhomogeneity fac- tor is small, the larger Pt atoms should be repelled from the interface on the Ni-rich side of the phase diagram and the smaller Ni atoms should be attracted to a GB on the Pt-rich side. This is inconsistent with the simulations, and, therefore, doesn’t explain the results. Alternatively, the explanation in terms of the inhomogeneity effect seems to work for the Pt-Au system, where the size misfit parameter is small and the inhomogeneity factor is large, that is, the “softer” Au atoms tend to be attracted to a GB on both sides of the phase diagram. The exact relation, however, between the magnitudes of the two elastic effects is determined by the value of an unknown dimension- less parameter, q.

The above discussion demonstrates that at present we cannot formulate a reasonably simple criterion to predict even the sign of the interfacial segregation! The connection between a bulk phase diagram of an alloy and interfacial phase diagrams for the same system is an intriguing and high priority area of research that can be resolved only on the basis of more detailed and accurate information than we presently have available.

3.4.2. Structural features. One macroscopic geo- metric DOF-the twist angle @--was systematically

Page 19: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1054 D. N. SEIDMAN et al.

varied for the (002) symmetrical twist boundaries investigated. We find (see Fig. 2) that Fwlvte steadily increases with increasing f3 until saturation occurs at about 22.6’ on both sides of the phase diagram. The effect is, however, a factor of two larger on the Ni-rich side. The widths of the segregation profiles normal to the interface decrease systematically with increasing 6. This is indicative of an elastic interaction between a GB and a solute atom-at least at low angles. The volume expansions at the GBs scale with the amount of solute-atom segregation.

The solute concentrations at different sites near the boundaries are found to be very inhomogeneous both for low- and high-angle boundaries. At low-angle boundaries some sites have concentratios of solute atoms significantly greater than in the bulk, that is, these sites are enhanced in solute, while other sites are unaffected by solute-atom segregation. At high-angle boundaries a considerable fraction of the sites have a concentration that is less than the bulk concentration, while other sites are very strongly enhanced. A more detailed discussion of solute-atom segregation at high-angle twist boundaries is given elsewhere [99].

The spatial distribution of sites affected by the solut~atom segregation at low-angle boundaries is found to be complementary on both sides of the phase diagram. On the P&rich side the enhanced sites are arranged in rows, above and below the cores of the PGBDs, in the form of an hourglass-like pattern based on the square grid of screw dislocations. Alter- natively, on the Ni-rich side the enhanced sites are arranged in the shape of a stepped bipyramid that is based on the same square grid of PGBDs. The sites on the outer surfaces of each bipyramid are more strongly enhanced in solute atoms than sites inside them. The hourglass-like structures account for the nomnonotonic segregation profile of the 0 = 5” boundary in Pt-3 at.% Ni, as the average concen- trations in the second (002) plane are higher than at the first (002) plane. In both cases the concentration of solute atoms at enhanced sites decreases with increasing distance from the interface. The spatial extent of the two geometric arrangements of solute atoms--hourglass-like structures and stepped bipyra- mids-decreases with increasing 6; this effect points to the importance of elastic interactions as the pen- etration depth of the elastic stress field associated with twist boundaries is inversely proportional to the periodicity of the dislocation grid by St Venant’s principle.

The results presented in this paper show abrupt changes in the solute-atom segregation behavior near the cores of the primary GB screw dislocations. On the Ni-rich side there is an abrupt transition from the highest level of enhancement at the surface of stepped

bipyramids to no enhancement at all beyond these surfaces. Alternatively, on the Pt-rich side the hour- glass-like structures above the dislocation cores are strongly enhanced in Ni but the effect disappears abruptly when crossing into the bipyramidal region. One possible explanation may be that the solute- atom segregation behavior is completely different in a dislocation’s core from its behavior in the elastic region beyond the core. This possibility appears unlikely, however, because the abrupt transitions occur in the 2nd, 3rd and 4th planes from the (002) geometric interface, and this is well beyond the possible radius of influence of a dislocation core. Another possible explanation is based on a compe- tition between the inhomogeneity and size misfit effects as these effects have different laws of decay. When these effects have different signs and are of comparable magnitude, the areas with the strongest interaction energies can be shifted from the dislo- cation cores. We cannot make any better prediction, because only an order of magnitude estimate for the dimensionless parameter, q, in this model is presently available.

The same low-angle twist boundaries in dilute Au-Pt alloys exhibit a stepped bipyramidal pattern of segregation on both sides of the phase diagram, that is, the bipyramidal structures are enhanced in Au on the Pt-rich side and are depleted in Pt on the Au-rich side. The remaining hourglass-like structures above the dislocation lines are virtually unaffected by solute-atom segregation.

Bipyramidal patterns of solute-atom segregation are also observed in MC simulations of (001) semico- herent heterophase boundaries in binary Cu-Au [ 1001 and ternary Cu-Ag-Au [loll alloys. The effect is larger in the latter case, where pyramids of a silver- rich phase form that penetrate deeply in the copper- rich phase. The bipyramids are based on the cells of a square grid of the geometrically necessary misfit dislocations.

A direct correlation between the local hydrostatic pressure and chemical composition was observed in Ag monolayers deposited on Cu substrates. A tight-binding molecular dynamics study [102, 1031 showed that the larger Ag atoms tend to occupy atomic sites in tension, while the smaller Cu atoms preferentially go to compressed sites forming a hexag- onal grid.

Recently, the effect of local stresses on solute-atom segregation at (002) boundaries was investigated for a Cu-5 at.% Ni alloy by the free-energy minimization technique in the local-harmonic approximation [104]. It was found that enhancement in Cu, which has a slightly larger lattice parameter, than Ni-, = 0.0233 for Ni(Cu) [105+orrelates strongly with tensile

Page 20: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Solute-atom segregation at grain boundaries 1055

stresses and is not affected by shear stresses. On the other hand, at a low-angle boundary in the same alloy substanti~ solute--atom segregation was ob- served in the 2nd and 3rd planes from the interface with a bipyramidal pattern. From the point of view of elasticity theory, there can be no tensile stresses there.

From the above we conclude that the conventional wisdom [106] about the atomic sites near the interface being essentially of two types-“good fit” ones where there is almost no distortion of the crystalline lattice and “bad fit” ones where distortions are large (in the cores of GB dislocations)--is inadequate to describe the atomic-scale picture of solute-atom segregation. In our case some “bad fit” sites are strongly affected by solute-atom segregation, while others are not. Accurate calculations of local stresses at each site can be very helpful in understanding interfacial segre- gation phenomena. These results are to be presented elsewhere.

1. Monte Carlo simulations were performed of solute-atom segregation at symmetrical (002) twist boundaries, as a function of the twist angle, for dilute single-phase alloys on both sides of the Ni-Pt phase diagram at 850 K.

2. The Monte Carlo simulations exhibit solute- atom enhancement at twist boundaries for both the Pt-3 at.% Ni and Ni-3 at.% Pt alloys. This result is in qualitative agreement with solute-atom segre- gation at free surfaces in the same alloy system (see discussion in [72] and references therein). The magni- tude of the enhancement for the grain boundaries studied is approximately a factor of two greater on the Ni-rich side as measured by the Gibbsian inter- facial excess of solute.

3. In both alloys the value of the Gibbsian inter- facial excess of solute increases with increasing 5 and it saturates at approximately one-half of the 45” interval, that is, at about the angle at which the cores of the primary grain boundary dislocations begin to overlap.

4. The widths of the solute-atom segregation profiles normal to the (002) geometric interface narrow with increasing 6 from three to two planes on each side of an interface until saturation occurs. This is consistent with the fact that the penetration depth of the elastic stress fields associated with a twist boundary decreases with increasing 5. At the smallest value of 5 studied (S’), on the Pt-rich side of the phase diagram, an oscillatory segregation profile is ob- served that disappears with increasing 5. On the Ni-rich side oscillatory segregation profiles are not observed.

5. At low-angle boundaries on the Pt-rich side of the phase diagram solute atoms tend to segregate in hour~ass-like structures based on the square grid of primary grain boundary screw dislocations. For the same boundaries on the N&rich side a spatially- complementary type of segregation pattern is ob- served, with the regions enhanced in solute atoms forming a stepped bipyramidal pattern that is cen- tered on the same square grid of primary grain boundary screw dislocations.

6. The height of the stepped bipyramidal and hourglass regions-that is, the number of (002) planes where soluteatom segregation is observed- decreases with increasing 5 from about five atomic planes each side of the interface to two atomic planes in the high-angle regime; in the latter regime the patches of perfect crystal become comparable to the areas of dislocation cores and their intersections. This is consistent with elasticity theory considerations [98].

7. A comparison with the previously studied Au-Pt system, where a reversal of the sign of the Gibbsian interfacial excess of solute occurs when changing sides of the phase diagram, demonstrates that the bulk thermodynamics of an alloy as well as the structure of a grain boundary plays an important role in determining the detailed atomic-scale behavior of soluteatom segregation.

8. A comparison with other available simulation results on soluteatom segregation at grain bound- aries, heterophase boundaries and free surfaces suggests that hourglass and bipyramidal patterns of solute-atom segregation are quite common, and are related to the stress-field pattern associated with interfaces.

9. No significant differences with respect to solute- atom segregation are observed between the so-called “special” high coincidence boundaries, as opposed to “general” boundaries. This contradicts common ex- pectations that “special” boundaries, due to a higher degree of order, should have weaker solute-atom segregation. The “special” boundaries are known to possess peculiarities in interfacial tensions and kinetic properties, but equilibrium segregation is determined by the partial derivative of the interfacial free energy of an interface with respect to the chemical potential of solute atoms [eqn (2)], and not the absolute value of the interfacial free energy.

Acknow~cd~e~en~s-This research was suonorted bv the National &exe Foundation under grant -number DMR- DMR-9319074 (Dr B. MacDonald, grant officer). This work is also partially supported by grant number DMR920002N. and it utilizes the CRAY Y-University of Illinois at Urbana- Champaign. This research made use of MRL Central Facili- ties supported by the National Science Foundation, at the Materials Research Center of Northwestern University, under award No. DMR-9120521.

Page 21: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

1056 D. N. SEIDMAN et al.

REFERENCES

1. Hondros E. D. and Seah M. P., in Physical Metallurgy (Edited by R. W. Cahn and P. Haasen), Vol. 1, pp. 856-931. North-Holland, Amsterdam (1983).

2. Briant C. L., J. Phys. (Paris) 49, C5-3 (1988). ’ Briant C. L.. in Materials Interfaces (Edited bv D. 3.

4.

5.

6.

7. 8. 9.

10.

11.

12.

13. 14.

15.

16.

17.

18

19. 20.

21.

22.

23.

24.

25.

Wolf and S.’ Yip), pp. 463486 Chapman & Hall, London (1992). Foiles S. M. and Seidman D. N., in Materials Inter- faces (Edited by D. Wolf and S. Yip), pp. 497-515. Chapman & Hall, London (1992). Seidman D. N., in Materials Interfaces (Edited by D. Wolf and S. Yip), pp. 58-84. Chapman & Hall, London (1992). Baskes M. I., Hoagland R. G. and Needleman A., Mater. Sci. Engng A 159. 1 (1993): Yoo M. H.. Sass S. L., Fu C. L.,-Mills M. J., Dimiduk D. M: and George E. P., Acta Metall. Mater. 41, 987 (1992). Hondros E. D., Muter. Sci. Engng A 166, 1 (1993). Watanabe T., Muter. Sci. EngngA 166, 11 (1993). Lim L. C. and Watanabe T.. Acta Metall. Mater. 38. 2507 (1990). Grain-Boundary Structure and Kinetics (Edited by R. W. Balluff?). American Society for Metals, Metals Park, OH (1980). Interfacial Segregation (Edited by W. C. Johnson and J. M. Blakely). American Society for Metals, Metals Park, OH (1979). Martin G. and Perraillon B., in Grain-Boundarv Structure and Kinetics (Edited by R. W. Balltij, PP. 239-295. American Societv for Metals. Metals Park, OH (1980). Ballti R. W., Met. Trans. 13A, 2069 (1982). Balluffi R. W. and Sutton A. P., Interfaces in Crys- talline Materials. Oxford University Press, Oxford, to be published in (1995). BalhttIi R. W., in Znterfucial Segregution (Edited by W. C. Johnson and J. M. Blakely), pp: 193-236. American Society for Metals, Metals Park, OH (1979). _ Seidman D. N., in Materials Interfaces (Edited by D. Wolf and S. Yip), pp. 58-84. Chapman & Hall, London (1992). Fitzsimmons M. R. and Burke1 E., Phvs. Rev. B 47, 8436 (1993); Fitzsimmons M. R. and Sass S. L., Acta Metall. 37. 1009 (1989): Maiid I.. Bristowe P. D. and BallutIi R.’ W., Piys. Reu. B. 40,’ 2779 (1989). Hall E~L,I, Phy&(ParaF) 43, XX-239 (l982); Mtdler D. A., Batson P. E., Subramanian S., Sass S. L. and Silcox J., to appear in Mat. Res. Sot. Symp. Proc. (1994). Foiles S. M., Phys. Reu. B. 40, 11502 (1989). Najafabadi R., Wang H. Y., Srolovitz D. J. and LeSar R., Acta Metall. Mater. 39. 3071 (1991). Seki A., Seidman D. N., Oh Y. and Foiles S. M., Acta Metall. Mater. 39. 3167, 3179 (1991). Udler D. and Seidman D. N., Phys. Stat. Sol. (b) 172, 267 (1992); Mat. Res. Sot. Symp. Proc. 278, 223 (1992); Actu Metall. Muter., 1959 (1994); Mater. Sci. Forum, lSklS6, 189 (1994); to appear in Interface Sci. (1994). Gibbs J. W., Trans. Corm. Acad. ZZZ 108, 343 (1875-78); Gibbs J. W., Collected Works. Yale Univ. Press, New Haven, CI (1957). Cahn J. W., in Znterfacial Segregation (Edited by W. C. Johnson and J. M. Blakely), p. 3. American Society for Metals, Metals Park, Ohio (1979); J. Phys. (Paris) 43, C6-199 (1982). Shuttleworth R., Proc. Phys. Sot. A 63, 444 (1950); Herring C., in The Physics of Powder Metallurgy (Edited by W. E. Kingston), p. 143. McGraw-Hill, New York (1951); Mullins W. W., in Metal Surfaces,

26.

27.

28.

29.

30.

31.

32.

33.

34. 35.

36.

37.

38.

39.

40. 41.

42.

43.

44.

45.

46.

47.

48.

49.

50.

51.

p. 17. American Society for Metals, Metals Park, OH (1962); Cahn J. W., Actu Metnll. 28, 1333 (1980). Hu J. G., Ph.D. thesis, Northwestern University (1991): Krakauer B. W.. Ph.D. thesis. Northwestern University (1993). Udler D. and Seidman D. N., Phys. Stat. Sol. (b) (1992); Mater. Sci. Forum lSS-156, 189 (1994); Acta Metall. Mater., 1959 (1994); to appear in Interface Sci. (1994). Rice J. R. and Thomson R. M., Phil. Mug. 29, 73 (1973). Rice J. R. and Wang J. S., Muter. Sci. Engng A 107, 23 (1989). Wang J. S. and Anderson P. M., Acta Metall. Mater. 39, 779 (1991); Wang J. S., Scripta Metall. Mater. 25, 133 (1991). Sun Y. M., Beltz G. E. and Rice J. R., Mater. Sci. Engng A 170, 67 (1993). Liu C. T., White C. L. and Horton J. A., Acta Metall. 33, 213 (1985). Seidman D. N., Hu J. G., Kuo S.-M., Krakauer B. W., Oh Y. and Seki A., CON. Phys. (Paris) 51, Cl-47 (1990). Seidman D. N., Mat. Sci. Ena. A 137, 57 (1991). Folies S. M., Baskes M. I. andDaw M: S., Phys.‘Reu. B 33. 7983 (1986): Phvs. Rev. B. 37. 10378 (1988). Udler D. and Seidman D. N., Acta ketall. kczter.‘42, 1959 (1994). Udler D. and Seidman D. N., to appear in Interface Sci. (1994). Krakauer B. W. and Seidman D. N., Phys. Reu. B 48, 1547 (1993); Krakauer B. W., Ph.D. thesis, Northwest- ern University (1993). Bostel A., Blavette D., Menand A. and Sarrau J. M., Coil. Phys. (Paris) SO, C8-501 (1989); Blavette D., Deconihout D., Bostel A., Sarrau J. M., Bouet M. and Menand A., Rev. Sci. Znstrum. 64, 2911 (1993). Cahn J. W., 1. Phys. (Paris) 43, C6-199 (1982). Ferrence T. G. and Balluffii R. W.. Scripta Metall. 22, 1929 (1988). Leamy H. J., Gilmer G. H. and Jackson K. A., in Surface Physics of Materials (Edited by J. M. Blakely), p. -121. Academic Press, New York-(1975).; Weeks J. D.. in Ordering in Stronalv Fluctuating Condensed Mutter Systems (Edited by-T. Riste), p. 293. Plenum Press, New York (1980). Hsieh T. E. and Balluffi R. W., Actn Metull. 37, 2133 (H89). Herschitz R., Seidman D. N. and A. Brokman, J. Phys. (Paris) 46, 0%451 (1985); Herschitz R. and Seidman D. N., Acta Metull. 33, 1547 & 1565 (1985); Herschitz R. and Seidman D. N., J. Phys. (Paris) 49, CS-469 (1988). Sickafuss K. and Sass S. L., Scripta Metall. 35, 69 (1987); Lin C. H. and Sass S. L., Scripta Metull. 22, 735 (1988); Scripta Metall. 22, 1569 (1988). Michael J. R., Lin C. H. and Sass S. L., Scripta Metall. 22, 1121 (1988). Blakely J. M. and Thapliyal H. V., in Znterfaciul Segregation (Edited by W. C. Johnson and J. M. Blakely), p. 137. American Society for Metals, Metals Park, OH (1979). Blakely J. M., in Encyclopedia of Materials Science and Engineering (Edited by M. B. Bever), Vol. 7, p. 4962. Pergamon Press, Oxford (1986); Blakely J. M., J. Phys. (Paris) 49, CS-351 (1988). Lagues M. and Domange J. L., Sur$ Sci. 47, 77 (1975). Senhaji A., Treglia G., Legrand B., Barrett N. T., Guillot C. and Villette B., Surf. Sci. 274, 297 (1992). Legrand B., Saul A. and Treglia G., Mater. Sci. Forum 154-156 (1994).

Page 22: ATOMIC SCALE STUDIES OF SOLVE-ATOM SEGREGATION AT …arc.nucapt.northwestern.edu/refbase/files/Seidman-1994_3074.pdfSolute-atom segregation at grain boundaries 1037 centrated on studying

Solute-atom segregation at grain boundaries 1057

52.

53. 54.

55.

56.

57.

58,

59.

60.

61.

62,

Barrett N. T., Villette B., Senhaji A., Guillot C., Belkhou R., Treglia G. and Legrand B., Surf. Sci. 286, 150 (1993). Smith C. S., Trans. AIME 175, 15 (1948). Gleiter H. and Chalmers B., in Progress in Materials Science (Edited by B. Chalmers, J. W. Christian and T. B. Massalski), Vol. 16, pp. l-272. Pergamon, Oxford (1972). Seidman D. N., in Materials Interfaces (Edited by D. Wolf and S. Yip), Chap. 2, pp. 58-84. Chapman & Hall, London (1992). Herschitz R. and Seidman D. N., Acta Metall. Mater. 33, 1547 (1985); Herschitz R. and Seidman D. N., Acta Metall. Mater. 33, 1565 (1985). Seidman D. N., Hu J. G., Kuo S. M., Krakauer B. W., Oh Y. and Seki A., .f. Phys. (Paris) 51, Cl-47 (1990). Krakauer B. W., Hu J. G., Kuo S.-M., Mahck R. L., Seki A., Seidman D. N., Baker J. P. and Loyd R. J., Rev. Sci. Instrum. 61.3390 (1990): Krakauer B. W. and Seidman D. N., Re;. Sci. kstr&n. 63, 407 1 (1992). Hu J. G. and Seidman D. N., Phys. Rev. Left. 65, 199 (1990); Hu J. G., Ph.D. thesis, Northwestern Univer- sity, 1991; Hu J. G. and Seidman D. N., Scripfa Metall. Mater. 27, 693 (1992). Krakauer B. W. and Seidman D. N., Phys. Rev. B 48, 6724 (1993); Krakauer B. W., Ph.D. thesis, Northwest- ern University, 1993. Lupis C. H. P., Chemical Thermodynamics, Chap. 14, pp. 391-399. Elsevier Science, New York (1983). Hu J. G. and Seidman D. N., Phys. Rev, Left. 65,199 (1990).

63. Krakauer B. W., Hu J. G., Kuo S.-M., Mallick R. L., Seki A., Seidman D. N., Baker J. P. and Loyd R., Rev. Sci. Instrum. 61, 3390 (1990).

64. Krakauer B. W. and Seidman D. N., Rev. Sci. Instrum. 63, 4071 (1992).

65. Udler D., Hu J. G., Kuo S.-M., Seki A., Krakauer B. W. and Seidman D. N., Scripta Metall. Mater. 25, 841 (1991).

66. Rittner J. D., Udler D., Seidman D. N. and Oh Y., submitted for publication (1994).

67. Udler D. and Seidman D. N., Acta Metall. Mater. 42, 1959 (1994).

68. Dahmani C. E., Cadeville M. S., Sanchez J. M. and Moran-Lopez J. L., .Pfrys. Rev. Lett. SS, 1208 (1985).

69. Gauthier Y., Joly Y., Baudoing R. and Lundgren J., Phys. Rev. B. 31, 6216 (1985).

70. Kuo S.-M., Seki A., Oh Y. and Seidman D. N., Phys. Rev. Lett. 65, 199 (1990).

71. Foiles S. M., in Surface Segregation and Related Pheno~na (Edited by P. A. Dowben and A. Miller), p. 79. CRC Press, Boca Raton, FL (1990).

72. Daw M. S., Foiles S. M. and Baskes M. T., Mater. Sci. Repts 9, 251 (1993).

73. Metropolis N., Rosenbluth M. N., Rosenbluth A. W., Teller A. H. and Teller E. J. Chem. Phys. 21, 1087 (1953).

74. Daw M. S. and Baskes M. I., Phys. Rev. Left. 50,128s (1983).

75. Foiles S. M. and Seidman D. N., in Materials Znter- faces: Atomic-Level Structure and Properties (Edited by D. Wolf and S. Yip), p. 463. Chapman & Hall, London (1992).

76. Daw M. S. and Baskes M. I., Phys. Rev. Lett. Jo,1285 (1983).

77. Bristowe P. D., Majid I., Counterman C., Wang D. and ~~ R. W., Mater. Sci. Forum 126-128, 25 (1993).

78. Wang H. Y., Najafabadi R., Srolovitz D. J. and LeSar R., Phil. Mug. A 65, 625 (1992).

79. Ashby M. F., Spaepen F. and Williams S., Acta Metali. 26, 1647 (1978).

80. Hirth J. P. and Camahan B., Actu Metall. 40, 1237 (1992).

81. Bristowe P. D. and Cracker A. G., Phil. Mug. A 38, 481 f 19781.

82. 83.

84.

85.

86.

87.

88.

89.

90.

91.

92.

93. 94.

95.

96.

97.

98.

99.

100.

101.

102.

103.

104.

105. 106.

Wolf D., Acta Metall. 32, 735 (1984). Evangelakis G. A. and Pontikis V., Europhysics Let?. 8, 599 (1989). Wang H. Y., Najafabadi R., Srolovitz D. J. and LeSar R., &erface Science 1, 31 (1993). Read W. T.. Dislocations in Crwtals. McGraw-Hill. New York (1953). Schober T. and Balluffi R. ‘UT,, Phil. Mug. 21, 109 (1970). Schwartz D., Vitek V. and Sutton A. P., Phil. Mug. 51, 499 (1985). Schwartz D., Vitek V. and Bristowe P. D., Actu MetalL 36, 675 (1988). Najafabadi R., Srolovitz D. J. and LeSar R., J. Mater. Res. 5, 2663 (1990). Seki A., Seidman D. N., Oh Y. and Foiles S. M., Acta Metall. Mater. 3P,3167 (1991); Acta Metail. Mater. 39, 3179 (1991). Udder D. and Seidman D. N., submitted to Phys. Rev. L&r. (1994). Udler D. and Seidman D. N., Phys. Stat. Sol. (b) 172, 267 (1992). Folies S. M., Phys. Rev. B 40, 11502 (1989). Najafabadi R., Srolovitz D. J., Wang H. Y. and LeSar R., Acta Metall. 39, 3071 (1991). Wang H. Y., Najafabadi R., Sroiovitz D. J. and LeSar R., Interface Science 1, 31 (1993). Bokshteyn B. S., Klinger L. M., Nikolsky G. S., Fradkov V. Y. and Shvindlerman L. S., Phvs. Met. Metullograph. 48, 75 (1981). Udler D. and Seidman D. N.. submitted for nubh- cation (1994). Udler D. and Seidman D. N., Scripta Metal& Mater. 26, 449 (1992). Udler D. and Seidman D. N., to appear in Interface Science (1994). Bather P., Wynblatt P. and Foiles S. M., Actu Metall. Mater. 39, 2881 (1991). Rogers III J. P., Wynblatt P., Foiles S. M. and Baskes M. I., Acta MetaN. Mater. 38, 177 (1989). TrCglia G., Legrand B., Eugene, Aufray B. and Cabani F., Phys. Rev. B 44, 6842 (1991). Mottet C., Treglia G. and Legrand B., Phys. Rev. B 46, 16018 (1992). Wang H. Y., Najafabadi R., Srolovitz D. J. and LeSar R., Acra Metall. Mater. 41, 2553 (1993). King H. W., J. Mater. Sci. 1, 79 (1966). McLean D., Grain Boundaries in Metals. Clarendon Press, Oxford (1957).