Anda di halaman 1dari 34

J. Phys. D: Appl. Phys. 30 (1997) 30373070.

Printed in the UK

PII: S0022-3727(97)69438-3

TOPICAL REVIEW

Surface effects of ordering in binary


alloys
M A Vasiliev
Institute of Metal Physics, National Ukrainian Academy of Sciences,
Vernadsky Street 36, 252680 Kiev, Ukraine
Received 7 February 1997, in final form 6 August 1997
Abstract. This article reviews the most important achievements in the studies of
the compositional (chemical) and magnetic ordering effects in the surface region of
single-crystal binary alloys with different bulk structures. These alloys include the
non-magnetic and magnetic concentrated substitutional binary systems which have
a tendency to ordering (AuCu, CuPt, FePt, CoFe, FeNi, AlFe, AlNi, PtTi,
CuPd, CoPt, NiPt, AlCu, CuAl). A number of analytical methods very
sensitive to the surface region have been recently developed to obtain a more
detailed knowledge of the crystallographic symmetry and lattice parameters, and
these as well as compositional and spin order parameters of the outermost few
atom layers of alloys are briefly described. Of primary interest are the following
phenomena observed from the experimental studies of the various ordering alloys:
the face-dependent bulk termination; surface composition and surface segregation
oscillatory profile; face-related sandwich segregation; strong surface multilayer and
rippled relaxation as well as buckling in the first few layers; surface reconstruction
and formation of the surface superstructure; the temperature-dependent surface
compositional and magnetic order parameters; kinetics of the orderdisorder phase
transition; surface electron structure; surface ferromagnetic anomalies. After a
review of the experimental data the interplay between theory and experiment is
also discussed. This comparison has proved particularly useful for understanding
the nature of the surface ordering effects in alloys.

6.1.

Contents
1.
2.

3.

4.

5.

6.

Contents
Introduction
Non-magnetic ordered solid solution
2.1.
Cu3 Au
2.2.
Cu3 Pt(111)
2.3.
Pt80 Fe20
Ferromagnetic ordered solid solution
3.1.
Co0.5 Fe0.5 : surface structure and
composition
3.2.
Ni3 Fe
Intermetallic ordered compounds: surface
layer structure and composition
4.1.
Fe3 Al(110)
4.2.
Ni3 Al
4.3.
Ni0.5 Al0.5
4.4.
Pt3 Ti(100) and (111)
4.5.
Pt3 Sn
Random solid solution
5.1.
Ordering solid solution
5.2.
Compound-forming alloys
Comparison of theoretical and experimental
studies

3037
3038
3039
3039
3046
3047
3048
3048
3049
3051
3051
3053
3053
3055
3055
3056
3056
3060
3061

c 1997 IOP Publishing Ltd


0022-3727/97/223037+34$19.50

7.

The surface orderdisorder phenomena


6.2.
Ordering and surface segregation
6.3.
Surface magnetism
6.4.
Surface effects in magnetic ordered
alloys
Conclusions
Acknowledgments
References

3061
3062
3064
3065
3066
3067
3067

1. Introduction
First it should be stressed that in the present review we
consider only the two types of ordering which take place in
the metallic alloys, namely, compositional (or chemical)
and magnetic. In the following, let us term these notions
for simplicity also order (ordering) and magnetization,
respectively.
The phenomenon of ordering in solid state physics is
well known [1, 2]. At the beginning we will note only
some general points in this connection. So, in the case
of a completely ordered substitutional alloy with specific
stoichiometric composition atoms of different sorts occupy
only a certain type of site in the crystal lattice. When the
3037

M A Vasiliev

temperature increases the atoms begin to fill foreign sites


and the concentration of the given sort of atoms on foreign
sites increases with temperature, but the concentration on
their own sites decreases.
At a certain critical temperature T0 in a completely
ordered alloy its symmetry changes during an order
disorder transition. As this takes place, the probability
of site substitution at T0 may change discontinuously (the
case of a first-order phase transition) or continuously (the
case of a second-order phase transition) depending on the
type of alloy. The ordering of atoms in the crystal lattice
may be characterized by the extent to which the different
sites forming a sublattice are occupied by the different sorts
of atom. Consequently, the ordering may be considered
in relation to the lattice sites, and the degree of order ,
called the degree of long-range order (LRO), is determined
by the arrangement of atoms over the entire crystal. The
degree of LRO from = 1 at room temperature decreases
with increase in temperature and disappears ( = 0) at the
critical temperature T0 .
It is well established that the ordering in alloys has
a great influence on their properties. So, the electric,
mechanical and other bulk properties of the alloys change
strongly upon ordering.
Recent interest in ordered
systems for high-temperature, high-ductility applications
has presented the challenge of understanding the geometric
and electronic structure of ordered binary alloys and
compounds. In the case of ferromagnetic ordered alloys
their properties are dependent on not only the ordering but
the magnetization which is characterized by mean atomic
magnetic moment [3]. The nature of the interplay
between ordering and magnetization in the bulk crystal is
well recognized [1, 4, 5].
However, the surface structure and properties of the
both non-magnetic and magnetic ordered alloys do not
necessarily reflect that of the bulk material. This may
be particularly important when considering, for example,
an alloys strength at a grain boundary or in detailed
understanding of its catalytic and corrosion behaviour, as
well as other chemical properties. In addition, adsorption
and desorption rates and chemical reactions on the surface
can be influenced by orderdisorder and magnetic phase
transitions. It is interesting that some ordering bimetallic
catalysts are known to have superior catalytic properties
for a number of catalytic reactions and are more active
than their pure components. However, ordered alloy
surfaces have received increasing attention in the last
few years not only in connection with their practical
importance, but also because of fundamental interest in
the field of the phase transition mechanism in solids: in
particular, the interrelation between the bulk and surface
phase transition behaviour. It is known that termination
of a solid creates a surface, which universally can change
its atomic arrangement in both the 2D surface lattice
and underlying atomic layers. Since the coordination of
atoms at the surface apparently differs from that of the
bulk, the conduction-electron distribution at the surface
is distinct from that in the bulk. As a result, the
electronic forces produced by this distribution cause the
new equilibrium crystallography structure of the surface to
3038

be different from that of an ideal truncated bulk surface


[6, 7]. As an example, a class of surface structure
changes, which have been studied extensively over the
past few years, are relaxation and reconstruction at the
clean surface of some pure metals, i.e. rigid movement
of one, or more, of the outermost surface layers with or
without any atomic rearrangement (reconstruction) within
the layers. The latter surface phenomenon is associated
with a specific surface phase transition in the pure metals
(W, Mo, Pt, Ir, Au). The studies of the layer relaxation
and reconstruction for monoatomic clean surfaces have
led to insights which motivated us to investigate surface
behaviour of the ordered single-crystal alloys. Thus,
obtaining detailed crystallographic informations for such
objects is a logical extension of past works on polyatomic
surfaces, since it can be inferred that various types of
relaxation and reconstruction could be present in binary
alloy surfaces. However, alloys have an extra degree of
freedom in the surface layers and can exhibit a number
of new phenomena which are absent in the bulk of the
crystal: in particular, different types of bulk termination
depending on face orientation; buckling of the surface
layers, i.e. a different relaxation of the sublattices; surface
segregation, i.e. enrichment of the free surface by one of the
components; surface reconstruction induced by the surface
segregation or adsorption of foreign species; surface order
in the absence of a LRO in the bulk and vice versa; changes
of the surface phase transition order, kinetics, and the
temperature-dependent LRO and magnetization parameters
in comparison with the bulk behaviour; last, a surface can
induce critical phenomena (wetting) in the alloys with a
bulk first-order transition.
All the noted specific surface phenomena in the case
of ordered (ordering) alloy systems have been predicted
from several theories since the first work by Valenta
and Sukiennicki in 1966 [8] and confirmed by suitable
experimental techniques since the first experimental studies
by Nielsen in 1973 [9] and Sundaram et al [10] in 1973.
Such experimental results and theoretical data are discussed
in the present review by the examples of a number of
concentrated substitutional alloy systems (non-magnetic
or magnetic) which have a tendency to compositional
ordering. It should be stressed that we have focused our
attention only on the surface properties of well defined
single crystals. Polycrystalline and amorphous alloys
are out of the scope of the review, because the data
obtained on the single crystal give, firstly, the fundamental
face-dependent characteristics for the clean free surfaces,
and, secondly, appropriate analysis in terms of basisdeveloped theoretical models. From the point of view
of an experimentalist who enjoyed surface research for
many years we consider here only those theoretical studies
which were compared with appropriate experimental data.
Also we consider only the experiments which have been
performed in conditions of ultra-high vacuum (UHV),
with in situ surface cleaning by cycles of noble gas ion
bombardment and thermal treatment in order to bring
the surface region to the equilibrium state. Let us list
the basic surface sensitivity techniques which have been
used to study the surface crystallographic, compositional,

Surface effects

electronic states and magnetization of the ordered systems


(in alphabetical order): Auger electron spectroscopy
(AES), ionization spectroscopy (IS), low-energy electron
diffraction (LEED), low-energy ion scattering (LEIS) with
time-of-flight (TOF) analysis, photo-electron spectroscopy
(PES), reflexion high-energy electron diffraction (RHEED),
scanning tunnelling microscopy (STM), spin-polarized
LEED (SPLEED) and x-ray diffraction (XRD). The more
comprehensive data one can find in the recent reviews
[11, 12] or in the appropriate references in the present paper.
The ordered (ordering) binary systems described in the
review are separated into two types: the concentrated solid
solution and compound-forming alloys, both non-magnetic
and ferromagnetic.
2. Non-magnetic ordered solid solution
2.1. Cu3 Au
The AuCu alloy system became the favourite object
for experimental as well as theoretical studies of the
relationship between bulk and surface ordering behaviour
mainly because its bulk properties and structure have been
well established [1, 2].
2.1.1.
Fcc L12 bulk structure. According to the
bulk phase diagram the AuCu alloys reveal two typical
compositional ordered phases: L12 (Cu3 Au or Au3 Cu) type
and L10 (AuCu) type [13]. In the disordered state such
alloys show the existence of a continuous series of solid
solutions of the substitutional type below the solidification
range. In this case the alloys have an fcc structure, and
atoms Cu (or Au) are encountered with equal probability
at all lattice sites. The alloys which are close to the
Cu3 Au composition are the most completely investigated.
The Cu3 Au phase is a classical ordering alloy, whose bulk
properties have been extensively studied. Such an alloy
with stoichiometric composition Cu3 Au becomes ordered
upon the attainment of an equilibrium state below a critical
temperature T0b of 663 K. In the ordered phase the Cu and
Au atoms rearrange in such a way that the Au atoms occupy
the corners of the cubic unit cell and the Cu atoms occupy
the face-centred sites (figure 1). Obviously, the number
of sites of the second type is three times greater than the
number of the first type. Thus the Cu3 Au fcc ordered lattice
can be subdivided into four simple cubic superlattices. This
superlattice structure can be determined, for example, by
x-ray diffraction methods which show superlattice Bragg
peaks. In this type of ordered alloy, as the temperature
increases the bulk order parameter = pAu pCu , the
difference of the probability of Cu sitting on corners
(pAu ) and on face-centred sites (pCu ), at first decreases
continuously to some non-zero value, and then jumps to
zero at the critical temperature. Thus a discontinuous first
orderdisorder transition occurs at T0b . It is important
to point out that the critical temperature decreases with
departure from the stochiometric composition in either
direction. The ordered phase exists over approximate
range 17 to 37 at.% of Au. This L12 ordered type
of superstructure is also exhibited by Au3 Cu, Cu3 Pd,

Figure 1. Unit cell for the Cu3 Au ordered bulk structure


(L12 superlattice).

Cu3 Pt, Fe3 Pt, Ni3 Fe, Ni3 Pt, Pd3 Fe, Pt3 Co, Pt3 Fe and
stoichiometric phases.
The AuCu alloys with fcc structure whose composition
is close to 50 at.% possess the CuAu L10 type of
crystal structure in the ordered state. In a completely
compositionally ordered alloy of stoichiometric CuAu
composition: both Au and Cu atoms are located in
alternating atomic planes. In this case, as in Cu3 Au the
degree of long-range order (LRO) at the orderdisorder
transition undergoes a step change. The disordered phase
exists at temperatures higher than approximately 685 K.
Such a type of superlattice is observed in CoPt, FePd, FePt
and also in some other alloys and compounds.
2.1.2. Surface structure and composition. Let us
consider the surface of only Cu3 Au among AuCu alloys
because it is the most completely investigated by several
surface sensitivity techniques. On the other hand the Cu3 Au
alloy showed the most definitive evidence for surface
ordering. Owing to the interest of the surface order
disorder transition, the fcc Cu3 Au alloy is the first one that
has been extensively studied above and below T0b .
(100) face. The first study of the Cu3 Au(100) face was
reported by Nielsen [9]. He has used RHEED technique and
observed the superlattice pattern that appeared to coincide
with the (100) plane of the bulk ordered Cu3 Au structure.
Sundaram et al [10] have performed the first qualitative
LEED study and a bulk termination periodicity was also
proposed for the Cu3 Au(100) face at T < T0b .
Other studies on the same alloy surface were performed
later in more detail also by quantitative LEED [14]. Potter
and Blakely [14] observed LEED patterns produced by
Ar ion bombardment at 773 K followed by annealing at
573 K and cooling to room temperature. This LEED
pattern was interpreted as c(2 2) superstructure expected
for the ideal termination of the ordered bulk crystal. It
means that the crystallographic structure of the outermost
layer is the same as that of a bulk crystallographic plane
of the same indices. However, in the case of ordered
Cu3 Au alloy, if a LEED pattern corresponding to bulk
termination was observed, that does not necessarily mean
that the surface superlattice is exactly established. This is
because in the bulk termination model the atomic planes
along a specific bulk crystallographic direction do not
necessarily have all the same composition. For example,
the atomic planes perpendicular to the h100i direction have
a stacking with a plane of composition CuAu alternating
3039

M A Vasiliev

with a plane of pure Cu. A priori mixed and pure Cu


surfaces correspond to bulk termination and in both cases
the compositions of the outermost planes are different from
the average one of the bulk (Cu0.75 Au0.25 ). The question
of the type of termination in the ordered Cu3 Au(100)
alloy was first answered by Buck et al [15] using LEIS
(TOF) supplemented by LEED and AES. It was found,
in agreement with Potter and Blakely [14], that LEED
superlattice spots indicated LRO as the sample was cooled
to room temperature after ion bombardment at 720 K.
The quantitative LEIS (TOF) results for room temperature
were consistent with an ordered (100) surface structure
like that of an ideally terminated fcc Cu3 Au(100) crystal
with Au atom fractions in the first and second surface
layers of 0.5 0.2 (mixed layer) and 0.00 (pure Cu layer),
respectively.
(110) face. This surface has been the subject of several
experimental studies [14, 1619]. It is interesting that the
(110) face, in contrast with the (100) one, did not exhibit
LRO according to the first qualitative LEED study by Potter
and Blakely [14]. They concluded that the Cu3 Au surface
exhibited a (1 1) diffraction pattern at room temperature
after annealing at 773 K. But after 20 h annealing at 573 K
a weak (2 1) type LEED pattern was obtained which was
the one expected of the ordered state. In the following
LEED study by Krummacher et al [16] it was shown that
the Cu3 Au(100) face yields a reconstruction with (4 1)
symmetry which was observed at room temperature after
surface cleaning by repeated cycles of Ar ion sputtering
and annealing up to 873 K. However at intermediate
temperatures around 473 K this symmetry transformed into
the (2 1) expected for an ordered, unreconstructed (110)
face. At a later time a more detailed analysis of both
structure and composition of the same face was performed
by McRae et al [1719] using LEED (PSD), AES and LEIS
(TOF) techniques. After a standard ion bombardment and
annealing at 625 K, the room-temperature LEED pattern
in this study indicated the (4 1) periodicity rather than
the (2 1) periodicity of the ordered alloy with ideal
termination. However the LEED patterns observed at
a different temperatures yielded a (4 1) (2 1)
transition at about 425 K. McRae et al [19] have performed
detailed LEIS composition observations on the (4 1) and
(2 1) surface symmetry. Since the ordered-bulk (110)
atom planes are alternately AuCu and pure Cu with regard
to composition there are two possible ideal terminations.
On the other hand for the completely disordered state the
first x1 and second x2 surface layer compositions are the
same, Cu0.75 Au0.25 . LEIS results indicated that at room
temperature the measured composition values more nearly
resemble those of the Au-rich ideal termination (x1Au = 0.5,
x2Au = 0) than of the Au-poor one. But there was a
substantial enrichment of Au in the outermost two atom
layers: the Au atom fraction in the first (second) atom layer
was found to change from 0.45 (0.20) at room temperature
to 0.35 (0.35) near T0b . These LEIS results corresponded to
a net segregation of Au atoms to the outermost two layers.
(111) face. The (111) face has not yet been thoroughly
investigated. Potter and Blakely [14] first performed
quantitative LEED analysis this face. They have noted that
3040

after annealing at 723 K and cooling to room temperature


the LEED pattern conformed with the disordered alloy, but
after annealing at 573 K for 20 h corresponded to the ideal
termination of the ordered alloy. In contrast to the (100)
and (110) surfaces in the ordered Cu3 Au each (111) plane
in the bulk is of stochiometric composition (Cu0.75 Au0.25 ).
However, Shaw and Fain [20] have found the surface of
Cu3 Au(111) to be enriched in Au (x1Au = 0.39).
2.1.3. LRO parameters and orderdisorder kinetics.
In recent years there has been quickened interest in the
study of the orderdisorder transition at Cu3 Au surfaces.
This classic ordering alloy which undergoes a firstorder transition provides an experimentally convenient
means of testing several theoretical speculations about the
relationship between bulk and surface ordering. Since
the average coordination in the surface region and the
interatomic electron distribution is changed in the outermost
layers one would expect surface ordering or disordering
behaviour somewhat different from those of the bulk.
Now the ordering or disordering behaviour has been well
investigated using a number techniques, including RHEED,
LEED, LEIS and x-ray diffraction methods. Those studies
yield important information on which ordered superlattices
occur, the kinetics of ordering and the variation of LRO
parameters and composition with temperature, particularly
near the critical temperature.
(100) face. The first evidence for the existence of the
orderdisorder transition at the (100) face of Cu3 Au was
reported by Nielsen [9] who used RHEED techniques. At
temperatures above 773 K the surface structure lattice unit
which
was found to be a square with a parameter of 2.65 A,
was in accordance with unit cell of the disordered structure.
At 643 K this structure transformed to the ordered state,
which corresponded to a structure with a square lattice
rotated 45 (with respect to the disordered one) with a
It was also confirmed that the
lattice parameter 3.75 A.
surface orderdisorder transition is reversible. Other studies
of the orderdisorder characteristics on the same face were
later performed by LEED, SPLEED and x-ray diffraction
[10, 17, 2124]. Sundaram et al [10] and Sundaram and
Robertson [25] have confirmed observation of ordering at
the (100) face of Cu3 Au, and in addition they have made
the first detailed study of the variation with temperature of
the LRO parameter s . To evaluate s as a function of
temperature the LEED beam intensity Ihk for a superlattice
was measured at different temperatures in the range 300
to 648 K, and in a single-scattering approximation the
following relation was used:
Ihk (s )2 e2M

(1)

where 2M is the well known DebyeWaller factor. Values


of the obtained s are shown in figure 2, as a function
of temperature. The bulk LRO b parameters determined
by x-ray measurement are also presented in figure 2.
One can see first from the figure that the surface order
parameter, in contrast to the bulk behaviour, appears
to be a continuous function of temperature. It is also
an important observation that the disordering process at

Surface effects

Figure 2. Temperature dependences of the measured bulk


() and surface () [10] LRO parameters from LEED of
Cu3 Au(100).

the surface begins at about 60 K below a common


critical temperature both for surface and bulk (T0b =
T0s ). Sundaram and Robertson [25] in addition to this
study investigated the temperature dependence of LEED
intensities in the range from 300 to 673 K and extracted
the effective surface Debye temperatures 2sD from the
data. For example, figure 3 shows the Debye plot of
log I /I0 versus temperature for the spicular (00) beam at
various electron energies E0 . It is interesting to note that
the anharmonic effects associated with the orderdisorder
transition are reflected in the non-linear behaviour of Debye
plots at temperatures above approximately 600 K. This
result supports the earlier conclusions [10] that the surface
disordering starts below the bulk value T0b . Indications of
a possible continuous orderdisorder transition on the same
face of Cu3 Au have been also recently reported [17, 21
24, 26]. McRae and Malic [21] characterized the surface
transition using LEED (PSD) with high angular resolution
and have found, in contrast to the abrupt change at T0b of the
x-ray superlattice, beam intensity decreases smoothly with
increasing temperature T in proportion to [(T0s T )/T0s ]2
with = 0.3. The same temperature dependence of
LEED intensity and beam profile shape was also observed
with temperature increasing or decreasing. The absence of
hysteresis in this temperature-dependent behaviour possibly
implied that the ordering time was small compared with
the experimental time constants. For example, the apparent
reversibility of LEED observations [21] which were made
with a time constant of the order of 10 s, implied that the
surface ordering was faster than the bulk ordering by a
factor of 105 .
SPLEED measurements performed by Jamison et al
[22] and Alvarado et al [23] also provide no evidence
of a sudden change in LRO at T0b = T0s . In addition to
this Alvarado et al [23] noted that a continuous surfaceinduced orderdisorder transition is equivalent to a critical
wetting phase transition [2729] on the Cu3 Au(100) face,
and indicated that near T0s the observed singularity was
related to the s which can be described by a functional
dependence of the form s t 1 , where t = 1 T0 /T0s
with 1 = 0.77 0.06. From all measurements they
obtained T0s = 662 15 K. The data also gave no evidence
of hysteresis. Thus the results obtained by LEED and
SPLEED have confirmed that the surface orderdisorder
transition on the Cu3 Au(100) face is of second order.

Figure 3. The logarithm of LEED intensity as a function of


temperature [25] for Cu3 Au(100).

Dosch et al [24], using x-ray diffraction under


total external reflection, have made the first attempt to
determine the order parameter profile as a function of
considering the Cu3 Au(100) surface.
the depth (>16 A),
The temperature dependence of the (100) evanescent
superlattice scattering intensity was measured in the
temperature range between 523 and 677 K (for a well
ordered surface at room temperature). Corresponding
temperature-dependent values of s are presented in
figure 4, and one can see that for small scattering depths
3 these data also exhibit the behaviour of a continuous
phase transition confirming results obtained by LEED and
SPLEED. On the other hand, if the values of 3 increases,
the features of a first-order (bulk type) transition can be
observed. Dosch et al [24] also found that the phase
transition temperature was independent of the distance from
the surface, while the surface layer started to disorder in the
presence of a still ordered bulk crystal. This situation was
characterized by an order parameter profile
t (z) = 1 [1 1s ] exp(z/b )
and
t (z) =

1/2 exp(z Lt )/b

1 1/2 exp(Lt z)/b

t > t

(2)

0 z Lt
t t

(3)

z Lt

where 1s is the continuous order parameter of the outermost


layer; b is the bulk correlation length; t = 1 T /T0b
the reduced temperature; t 4 102 is the point of
the strong change in the temperature dependence of t (3);
Lt is the mean position of the interface from the surface
(see figure 5). Finally Dosch et al [24] considered the
proportional relation between intensities It (3) and |t (3)|2
and concluded that the effective thickness of a disordered
surface layer which exists between the ordered bulk and the
vacuum at temperatures below T0b grows logarithmically as
the crystal temperature increases up to T0b .
(110) face. In observation of (110) face ordering
kinetics we refer to a surface that below T0b has the LEED
3041

M A Vasiliev

Figure 4. The temperature dependence of the surface


LRO parameter (x-ray) [24] of Cu3 Au(100).

Figure 6. Temperature dependences of the LEED intensity


for the beams indicated [19] for Cu3 Au(110).

Figure 5. LRO parameter profiles s (z )/b associated with


surface-induced disorder: (a) onset of surface disordering,
(b) advanced surface disordering [24].

pattern (2 1) symmetry, which is that expected for ideal


termination of the ordered bulk crystal. Krummacher et al
[16] using qualitative LEED showed that surface transition
from the ordered (2 1) symmetry to the disordered and
unreconstructed (1 1) structure occurred at the same
temperature T0s as in the bulk T0b = 663 K. A more detailed
analysis of the ordering kinetics was performed by McRae
and Buck [17], McRae et al [19] and McRae and Malic [18]
using LEED (PSD) and LEIS. It was found first [18, 19]
that the superlattice beam intensity decreases linearly with
increasing temperature and dies out at a temperature T0s
6 K below T0b . As shown in the LEED (PSD) intensity
versus T plots in figure 6 the transition exhibits marked
hysteresis associated with the (2 1) (1 1) transition
due to a relatively low rate of ordering. On the basis
of the asymmetry between ordering and disordering, the
surface transition was identified as first order, that is to
say a discontinuous surface orderdisorder transition. For
example, disordering at 651 K was an order of magnitude
3042

faster than ordering at the same temperature. The slow


variation of intensity near T0s , or the broadening of the
surface transition was expanded by analogy with the offstoichiometric bulk transition. This analogy is possible in
view of the observed surface segregation of Au (see below).
McRae et al [19] have also concluded that the observed
hysteresis and a symmetry of ordering and disordering rates
are expected since, on cooling, small ordered domains must
nucleate and grow in a disordered surface phase whereas on
heating disordered regions must form in an ordered phase.
Detailed observations of ordering kinetics were reported by
McRae and Buck [17], corresponding to two different initial
conditions of disorder: (i) disordered surface on an ordered
substrate; (ii) disordered surface on a disordered substrate.
It was noted that the times required to restore the surface
to a substantially ordered condition after annealing were
(i) 105 and (ii) 106 s. The latter value is comparable with
the bulk ordering time. On the basis of study of the ordering
kinetics at the Cu3 Au(100) face McRae and Malic [18] have
made a very important conclusion about an unpresented
case of two-dimensional compositional ordering, i.e. the
existence of ordering in two dimensions in the absence of
an ordered substrate to act as a template at some regime of
annealing. It is known [30] that according to the theoretical
interpretation in the case of discontinuous bulk transitions
disorder can be induced by a surface and propagates into
the bulk. Consequently it may be expected that ordering of
the crystal starts in the bulk and propagates to the surface
region. Put another way, the surface orders on top of an
already ordered substrate as on a template. The results on
the late stage of ordering obtained by McRae and Malic
[18] are in line with this assumption. However for the
earlier stage of ordering they found that the surface orders
first. Thus the kinetics for the Cu3 Au(110) face exhibit a
very complex picture.
(111) face. There is little information on ordering
kinetics at the Cu3 Au(111) face. Potter and Blakely [14]
first briefly pointed out the noticeable difference in the
ordering kinetics of the (111) face from that of (100), being
much faster on the (100) face. But the reason for this fact

Surface effects

Figure 7. Au concentration of the first and second layers of


Cu3 Au(100) as a function of annealing temperatures [15].

is not understood. McRae and Buck [17] have made the


conclusion that the surface transition at the (111) face has
properties similar to those for the (110) surface: a firstorder transition.
2.1.4. Temperature-dependent surface composition.
(100) face.
As was mentioned above the ordered
Cu3 Au(100) face at room temperature preserves practically
ideal bulk termination with outermost layers alternating
between mixed Cu0.5 Au0.5 and pure Cu. There is only a
little tendency of Au segregation (x1Au = 0.53, x1Au = 0.06).
The first data on the equilibrium surface composition as
a function of temperature were reported by Krummacher
et al [16]. They have made the conclusion that after
the surface cleaning procedure the intensity ratio of the
Au peak at 60 eV and the Cu peak at 70 eV in the
AES spectra showed no change in surface stoichiometry
in the temperature range from room temperature to 873 K.
However AES gives only the average layer composition.
The first temperature dependence of the outermost layer
composition was observed by Buck et al [15] using LEIS
(TOF) method with monolayer resolution. The variations of
the first- and second-surface-layer equilibrium composition
with the temperature are shown in figure 7. One can
see in this figure that the first-layer composition behaviour
yields clear evidence of the strong competition between
ordering/disordering and surface Au segregation. So,
for the first layer the Au concentration increases with
temperature, reaching the value x1Au = 0.62 at 673 K (close
to T0b ), and then monotonically decreases to approximately
its initial value 0.50 at 873 K. The latter value is rather
far from the average value of 0.25 for disordered alloy at
high temperature. This was the first observation of the
composition maximum in the temperature-dependent curve
x1 (T ) at near T0b . This effect will be discussed in the
theoretical part of the present review.
Now let us note that although the first layer does not
achieve the random bulk composition (xbAu = 0.25) in the
temperature range investigated, the second layer does (see
figure 7), but at a temperature above the T0b . This is because
of the AuCu bonding preference and the lattice strain
involved in placing more gold atoms in the second layer
while the first-layer concentration is so high [31]. Thus

the LEIS (TOF) study by Buck et al [15] gave evidence


that the compositional ordering tendency suppresses the Au
segregation and progressively disrupts the order, as shown
by the continuous weakening of the LEED superlattice
beams and LRO parameter [10]. On the other hand, the
surface Au segregation even for temperatures well above
T0b gives evidence of the existence of short-range order,
because the internal (antiferromagnetic nearest-neighbour)
interactions still favour CuAu interactions. One interesting
question arises from this observation: is the Au atom
concentration in the first and second outermost layers
really the type of oscillating depth profile which was
predicted theoretically [32, 33]? In order to answer this
question recently Reichert et al [34] have examined the
Cu3 Au(100) surface at temperature above T0b by means of
x-ray crystal truncation rod (CTR) diffraction. They first
have provided detailed information about the temperaturedependent surface segregation depth profile down to several
layers with atomic resolution.
The final results are
presented in figure 8 and figure 9. Figure 8 shows for
three selected temperatures, the layer-dependent excess
Au concentration 1xiAu = xiAu (T ) xbAu . One can see
here a sharply defined oscillatory behaviour that decays
exponentially versus depth with a decay length 3. This
value in units of a0 /2 (see figure 9) versus the reduced
temperature t = (T T0b )/T 0 in a double-logarithmic
plot exhibits a power-law behaviour with an exponent =
0.49 0.04. Thus even in layers at large distances (n = 0
10) from the first layer the Au composition can deviate from
its bulk value, and while the top-layer Au concentration
is high up to high temperatures, the decay length of the
oscillatory segregation profile decreases distinctly upon
heating. Interestingly, the decay length was found to scale
with T in the same way as the bulk correlation length in
mean-field theory. Hence the work by Reichert et al [34]
presented an example where the surface phenomenon is
apparently controlled by the bulk correlation length [35].
(110) face. Layer composition and related order
disorder transition at Cu3 Au(110) face were first described
in detail by McRae et al [19]. Using the combination
of LEED and LEIS (TOF) they revealed in contrast to
the (100) face, a pronounced dependence of surface layer
composition on temperature (figure 10). One can see
a substantial Au concentration in the second layer and
enrichment of Au in the two outermost layers compared
to the mixed ideal termination (x1Au = 0.5, x2Au = 0).
Figure 10 shows that with temperature increasing the Au
composition deviates further from room-temperature values
to an extent which grows fast near 425 K. Around T0b the
compositions have reached a level x1Au = x2Au 0.35,
which is moved closer to the ideal values x1Au = x2Au = 0.25
for the disordered state. It was shown previously that a
(21) (11) orderdisorder transition takes place near
the bulk disordering transition temperature T0b = 663 K.
McRae et al [19] give following original interpretation of
this surface transition behaviour considering the bulk phase
diagram and surface composition curve under heating (see
figure 11). They suggested an analogy with the two-phase
region associated with an off-stoichiometric discontinuous
first-order transition in the bulk of Cu3 Au alloy. Thus the
3043

M A Vasiliev

Figure 10. Au concentration of the first and second layers


of Cu3 Au(110) as function of annealing temperatures [19].

Figure 8. The decay of the Au layer concentration profile


for selected temperatures (1xiAu is the excess of the Au
concentration to the average bulk value 1xbAu [34]) for
Cu3 Au(100).
Figure 11. Schematic interpretation of the Cu3 Au bulk and
surface phase transition [19].

Figure 9. The double-logarithmic temperature dependence


of the decay length 3 for Cu3 Au(100) [34].

observed Au enrichment of the two outermost surface layers


exceeding the bulk stoichiometric value xbAu = 0.25 gives
rise to the orderdisorder transition outside the maximum of
the x(T ) plot (figure 11) corresponding to the bulk critical
temperature T0b .
2.1.5. Photoemission effects. Recently angle-resolved
UV photoelectron spectroscopy was demonstrated as the
most useful method for probing the wave-vector kdependence of the electronic structural density in metals
and alloys. It was also shown that this technique is very
suitable to study the electronic driving force of the bulk
and surface phase transition. In particular, we consider
3044

here the studies describing the electronic consequences of


the compositional orderdisorder transition at the Cu3 Au
surfaces [16, 3638].
(100) face. The first comparative study of the electron
structure of both ordered and disordered Cu3 Au(100)
face in the temperature range from room temperature
to 773 K was performed by Jordan et al [36], and
later re-examined by Jordan [37], using angle-resolved
photoelectron spectroscopy. An initial series of UV
photoemission measurements were carried out using HeI
photons (21.2 eV) in order to locate regions in Ek
space (0XRM plane for ordered and 0XLK plane in the
disordered state) where the effects of the orderdisorder
transition are expected to be the most evident in spectra.
The main changes were observed in the emission angles,
2, between 20 and 60 at different temperatures. The most
striking features were observed in the range 20 2
35 . Several features were found (figure 12) which varied
with temperature and were associated unambiguously with
orderdisorder transition. In fact two essential effects must
be pointed out: (i) when the alloy starts to order new
features associated with changes in the electronic structure
(bands) of the ordered phase appear in the spectral density
whose weights increase with the degree of order; (ii) the
strong temperature dependences of the features I and II
(figure 13) imply that their weights in the spectral density
increase at the same rate as a function of the order parameter

Surface effects

Figure 12. Photoemission spectra at different temperatures


for Cu3 Au(100) [37].
Figure 14. Comparison of valence band spectra of ordered
and disordered Cu3 Au(100) and Cu3 Au(110) surfaces at
(a) 26 eV and (b) 40 eV photon energy [16].

Figure 13. The temperature variation of features I and II


(figure 12) [37].

measured early on by LEED [10]. Since the electron


mean free paths in the photoemission measurements (h =
40 eV) are comparable with those in the LEED study,
those results give some information on the electronic
consequences of ordering at the surface region.
In the following year Krummacher et al [16] extended
the comparative study of the electron structure of the
ordered and disordered Cu3 Au(100) face using angleintegrated photoelectron spectra in the energy region
between 12 and 150 eV. The angle between the photon
beam and the surface normal was about 60 and an energy
analyser collected all electrons of the proper energy under
polar angles of 36 to 42 with respect to the analyser
axis. The spectra observed under these conditions provide
data on the surface electron density of states, because
the escape depth of the photoelectrons was always less
In this study analyses of experimental
than about 10 A.
photoelectron spectra were performed only for two fixed
temperatures: room temperature (ordered state) and about
700 K (disordered state). As an example figure 14
shows the comparisons of a sequence of valence band
spectra obtained at 26 and 40 eV photon energies for both
mentioned states. In the line with the interpretation by
Wertheim [39] region I (figure 14(a)) in the valence band
spectrum between about 7.5 eV and about 4.5 eV below

EF may be specified by Au 5d bands, and the region II


about 1.7 eV below EF by both Au 5d and Cu 3d bands.
Finally, the region from about 11.0 eV to 7.5 eV (not
marked in figure 14(a)) and from 1.7 eV below EF up
to EF (marked as region III) is dominated by sp-like bonds
of Cu and Au. There are also three maxima in region I
below EF marked as A, B and C, respectively. Let us turn
to comparison of the two types of valence band spectrum
(see figure 14(b)). First at all spectra for the disordered
state are in general more smoothed out as compared with
the ones for the ordered state. In addition, peak A does
not show any pronounced intensity maximum with varying
photon energy as in the case of the ordered surface. The
strong variations of the valence band with different surface
structure in region II, determined by van Hove singularities
and crystal field splitting, and strongly influenced by matrix
element and final-state effects, were not discussed in detail
[16].
(110) face. The first study of the comparison of valence
band spectra for both ordered and disordered Cu3 Au(110)
face was also performed by Krummacher et al [16]. The
photoelectron spectra of the ordered phase were obtained at
room temperature, and those of the disordered surface also
at the same temperature, but after heating and quenching.
The experimental conditions of the photoelectron spectrum
observation were similar to those described for the (100)
face [16]. Figure 14 shows the comparison of the valence
band spectra both for ordered and disordered states, as an
example, for 26 and 40 eV photon energy. One can see that
in general the behaviour of the corresponding spectra for
the disordered state is similar to that of the spectra of the
ordered, except for a slight broadening. The interpretation
of the main features in this case was also the same as was
noted above for the (100) face. Krummacher et al [16]
have suggested that the strong variations with different
surface phases in region II of the valence band spectra
are predominantly due to final-state effects. On the other
hand the region I was much less sensitive in this respect.
There are only differences in the relative intensities of peaks
3045

M A Vasiliev

A, B and C for the four surface structures and a slight


broadening for the disordered surfaces. This was surprising
since the (110) face exhibits a reconstruction whereas the
(100) face does not. At the other extreme observed by
Krummacher et al [16] was the absence of the ordering
and face orientation effects in the surface core level shift
of the Au 4f level. There was only a difference in relative
intensity probably because the (110) surface is more open
as compared with the (100) surface.
2.2. Cu3 Pt(111)
Cu and Pt atoms form a continuous series of fcc exothermic
disordered solid solutions at high temperatures and ordered
phases at low temperatures (Cu3 Pt, CuPt, Pt3 Cu and Pt7 Cu)
[13]. The CuPt alloys are interesting in the field of
catalysis because this system is catalytically more active
than pure Pt. This is a surprising result since Cu itself is less
active than Pt for oxidation and reforming reactions [40]
and has been explained controversially in terms of structural
and electronic effects. The Cu3 Pt(111) ordered surface
has been extensively studied for its adsorptive properties
[4042] since it allows the properties of isolated Pt atoms,
surrounded only by Cu atoms to be examined. On the
other hand, the CuPt system is suitable to test a theory of
surface segregation, which would account for lattice strain
caused by the difference in the atomic radii (about 8%,
and rP t = 1.39 A)
[43].
rCu = 1.27 A
The stoichiometric Cu3 Pt alloy has the same ordered
L12 structure as fcc Cu3 Au with Pt atoms at the corners
of the cubic unit cell and the Cu atoms at the face-centred
sites. The first study of the ordered Cu3 Pt(111) face was
performed by Schneider et al [40, 42] using LEED, AES,
UPS and work function measurements. The ordered state
was achieved by Ar ion sputtering and annealing just below
850 K. Surface ordering was confirmed by the LEED pattern which showed (22) superstructure with respect to the
(1 1)-(111)-fcc structure. Figure 15 shows a schematic
model proposed for the ideal surface structure. It may be
noted that the Pt atoms form a sublattice with the double
lattice distance of the fcc structure and that Pt atoms in
this ideal case are fully surrounded by six Cu atoms. A Pt
surface content of about 30 at.% was evaluated by AES.
This result is close to the ideal mixed composition of the
ordered Cu3 Pt(111) surface. Figure 16 shows the HeII UP
spectrum of the ordered Cu3 Pt(111) face in comparison to
the spectra of clean Pt(111) and Cu(111) in the valenceband range. The features observed at 0.9 and 4.0 eV and
the peak at 2.4 eV are contributions of Pt and Cu 3d states,
respectively. The intensity at the Fermi level is not sharp
and structured in contrast to Pt(111) surface. The Pt 5d
broadening seems to be decreased. Observed UPS results
of valence-band emission were interpreted as a change to
d10 state of the Pt atoms by alloying with Cu. The observed
work function of the ordered surface was 5.4 eV which is
approximately the stoichiometric superposition of the values of the two constituents (5.2 eV). The work functions of
Cu(111) and Pt(111) are 4.95 and 5.95 eV respectively [42].
In order to appreciate the meaning of the Pt
coordination on the adsorption properties of the CuPt alloy
3046

Figure 15. Schematic surface structure of ordered


Cu3 Pt(111) [40].

Figure 16. HeII UP valence-band spectrum of the ordered


Cu3 Pt(111) surface and clean Pt(111) and Cu(111) [40].

Schneider et al [42] and Castro et al [41] have studied Xe


and CO interaction not only with the ordered Cu3 Pt(111)
face, but with the disordered one as well. The preparation
method of the disordered surface consisted in heating the
sample at 760 K (after Ar ion bombardment) followed by
slow cooling. In this case the LEED pattern corresponded
to a simple (11)-(111)-fcc surface structure. These results
were consistent only with an indistinguishability of the two
kinds of surface atom. AES and work function observation
(5.65 eV) of the disordered surface allowed the estimation
of a surface composition of the outermost layer of at
least 60 at.% Pt. Consequently, the disordered Cu3 Pt(111)
surface is characterized by strong Pt atom segregation.
Castro et al [41] have also revealed the difference in
the interaction of CO with the Cu3 Pt(111) surface in the
temperature range 50400 K, using LEED, AES, TDS,
UPS and work function measurements. The conclusion was
supported that CO chemisorbs preferentially on the Pt sites
on both surfaces leaving Cu atoms free, but CO was more
strongly chemisorbed on the ordered surface than on the
disordered one.
In a recent study Shen et al [43] have performed
closer examination of the Cu3 Pt(111) alloy surface. They
used a combination of LEIS and LEED to determine the
composition of the first two outermost layers and to give
proof of the existence of the surface ordering effects in
real space. After Ar ion bombardment and long-time
annealing at about 850 K (near T0b ) the Cu3 Pt(111) surface
showed a sharp LEED (1 1) pattern which is typical of
a disordered and unreconstructed surface. However, the
detailed LEIS analyses have demonstrated the existence of
short-range order. LEIS also showed under equilibrium

Surface effects

conditions a tendency to surface segregation, namely, the


top layer had 80 at.% Cu20 at.% Pt, while the second layer
was 69 at.% Cu31 at.% Pt. This observed oscillation in
surface layer composition is typical as generally predicated
for alloys with a negative enthalpy of mixing (1Hmix =
10 kcal mol1 for Cu3 Pt). It was also indicated that the
Pt atoms in the first layer are not clustered in domains and
are coplanar with the Cu atoms.
2.3. Pt80 Fe20
The enhanced catalytic properties, as compared to pure Pt,
of alloys between Pt and Fe have indicated the studies of
their surface region concerning both surface structure and
segregation [4446]. FePt alloys show a strong tendency
to bulk compositional order. In the range 2583 at.% Pt
three ordered phases are formed based on the composition
Fe3 Pt (1 , T0b = 1108 K), FePt (2 , T0b = 1588 K) and
Pt3 Fe (3 , T0b = 1623 K) [13]. The ordered superlattice
phases Fe3 Pt and Pt3 Fe are both of the Cu3 Au (L12 ) type
of structure. The first phase is ferromagnetic and the second
is paramagnetic.
2.3.1. (111) face. Because the catalytic activity was
found to be strongly enhanced in a narrow composition
range, 80 5 at.% Pt, the Pt80 Fe20 alloy faces (111) and
(110) have been studied extensively. In the ordered state the
Pt80 Fe20 alloy is very similar to the stoichiometric 3 phase
(Pt3 Fe) and the excess Pt atoms (5% here) are randomly
distributed on the iron sublattice. The first quantitative
LEED analysis of the structure and composition in the
surface region for the ordered Pt80 Fe20 (111) alloy were
performed by Beccat et al [44]. The surface was cleaned
by repeated Ar ion bombardment followed by annealing at
1200 K. As a result for the clean (111) face, a (2 2)
LEED pattern (with respect to the disordered phase)
was observed. The I (V ) spectra for both and 3 phase
beams were calculated using two LEED programs based
on the layer-doubling method for stacking layers and the
combined-space method. The metric distances between
the experimental and theoretical I (V ) spectra were used to
optimize the following parameters: the interlayer spacings
di,i+1 ; the Pt concentration xiP t (i = 1, 2, 3); a buckling
in the outermost two layers dz1 and dz2 . It was first found
that the surface model including the compositional order
only, i.e. Pt3 Fe-like with no atomic displacements, did
not exhibit a satisfactory agreement between theory and
experiment. The best agreement was observed including a
monotonically decreasing Pt enrichment and also a bucking
of the pure Pt top layer, induced by compositional ordering
of the Fe sublattice in the second layer and in the bulk.
The final model of the surface region proposed for the
ordered alloy Pt80 Fe20 (111) is shown in figure 17. It was
found that the first surface layer was composed of almost
100% Pt (x1P t = 96 4 at.%) and the second layer was
a Pt3 Fe-like structure. It also appeared that these toplayer Pt atoms have two different possible environments
in the second layer (figure 17(a)): one atom (type 1) is
in contact with three Pt atoms of the second layer, the
others (type 2) being in contact with two Pt and one Fe.

(a)

(b)
Figure 17. The final model of Pt80 Fe20 (111) surface:
(a) surface reconstruction; the top layer contains
100 at.% Pt, the second one is Pt3 Fe-like; (b) side view,
cut along the AA line of (a) [44].

The Pt3 Fe-like second layer and different Pt and Fe radii


respectively) induce the following atom
(1.385 and 1.241 A,
movements: the top-layer Pt atoms ride up the Fe atoms in
the second layer, producing the buckling dz1 and in-plane
movement dh, such that type-2 atoms are repelled by Pt
atoms of layer 2 and attracted by Fe atoms (figure 17(b)).
For the optimum parameter and surface layer composition
it was found that x1P t = 96 4, x2P t = 88 7, x3P t =
85 15 at.%; 1d12 = +0.3 1.1%, 1d23 = 0.6 1.8%,
and dh 0.03 A.
Thus for the
1z1 = 0.09 0.02 A
ordered Pt80 Fe20 (111) alloy the composition profile was
found to be monotonically decreasing in the first three
outermost layers instead of oscillatory surface segregation
as for other ordering alloys with the same orientation (see
NiPt alloys).
2.3.2.
(110) face. Also the surface of an ordered
Pt80 Fe20 (110) single crystal was studied by BaudoingSavois et al [45] using quantitative LEED intensity
analysis. For the ideal Pt80 Fe20 (110) surface in the [110]
direction normal to the (110) face, the crystal exhibits as
known a stacking sequence of two different layers with
60 and 100 at.% Pt. As one would expect in this case,
two bulk ideal terminations are possible: a sequence of
layers containing 60, 100, 60, . . . at.% Pt, respectively, and
a sequence 100, 60, 100, . . . at.% Pt, respectively.
LEED analysis have shown that the (110) face
of the ordered Pt80 Fe20 crystal exhibits two different
superstructures described as (1 2) and (1 3)
3047

M A Vasiliev

reconstruction. Both superstructures were produced by


annealing the Ar-ion-bombarded surface: the (1 3) in the
range 750 to 900 K; and (1 2) above 1000 K. BaudoingSavois et al [45] have performed a detailed LEED analysis
of the (1 2) stable superstructure only. Because of the
striking similarity in the I (V ) curves of the pure Pt(110)
and Pt80 Fe20 (110)-(12) surface they chose the well known
missing row model with geometrical parameters of the pure
Pt(110)-(12) reconstructed structure as a start model. The
calculation procedure of the I (V ) intensity was the same
as in the early study [44]. Neglecting compositional order
16 parameters were optimized in the structure analysis,
including nine geometrical parameters describing the lattice
relaxations within the five outermost layers (figure 18(a)),
and seven additional parameters allowing a different Pt
concentration in each non-equivalent row.
In order
to consider an ordering effect ten additional parameters
describing the Pt site concentrations within the rows in
layers 1, 3, and 5 were analysed (in layers 2 and 4, the
four atoms/unit cell were symmetrically equivalent). Nonequivalent occupations of lattice sites also allowed different
geometrical parameters. Figure 18 shows the optimum
structure of the Pt80 Fe20 (110) surface region based on the
missing row (1 2) reconstruction. Optimum geometrical
(1d12 =
parameters were determined to be d12 = 1.2 A
d34 = 1.40 A,
d45 = 1.41 A,

13%), d23 = 1.40 A,


1z3 = 0.15 A,
1y2 = 0.03 A,
1y4 =
d56 = 1.40 A,

(db = 1.379 A).


In figure 18(b) a limited grey
0.01 A
scale represents different levels of Pt concentration (black,
grey, and white circles denote 100, 80 and 60 at.% Pt,
respectively). In connection with surface layer composition
and ordering several interesting facts were pointed out:
(i) the outermost visible row with only 13 at.% Fe
appears totally distorted; (ii) the ordering gradually recovers
the bulk situation over five to six layers; (iii) general
Pt enrichment is found in the surface visible rows (in
layers 13), but segregation and composition order yield
a subtle redistribution of Pt and Fe atoms in deeper rows,
for example, in layer 2, the visible row is Pt rich, whereas
the other row (buried under layer 1) is enriched with Fe.
Finally with missing row (1 2) structure the average Pt
concentrations are in the first layer 87, the second 60 and
the bulk layer 80 at.% Pt. This would correspond to a Pt
composition oscillation starting with Pt enrichment in the
surface. Finally we comment that the surface of the ordered
Pt80 Fe20 (110) crystal is example of very complex structure,
including reconstruction, segregation, atom displacements
and partial order. The missing row model for this surface
was also confirmed in a recent STM study by Hammar et al
[46], although different (11) phases were found to coexist
with the (2 1) structure.
3. Ferromagnetic ordered solid solution
3.1. Co0.5 Fe0.5 : surface structure and composition
The ferromagnetic CoFe alloys, from the viewpoint of
technical importance, possesses several useful characteristics, particularly such as the largest known magnetization
per atom and an extremely high Curie temperature. There
3048

(a)

(b)
Figure 18. Pt80 Fe20 (110) (1 2) missing row model:
(a) the parameters optimized in the structure analysis
(d1 d5 interlayer spacing; 1z3 , 1z5 buckling in odd layers,
1y2 , 1y4 lateral shift); (b) perspective view of the optimum
structure [45].

is also considerable current interest in studying the catalytic


properties of CoFe alloys, for example in the hydrogenation reactions.
The CoFe phase diagram [13] shows a wide range
of solid solution with bcc Fe and a region from 30 to
70 at.% Co where the ordered bcc Co0.5 Fe0.5 (B2, CsCl
type) superstructure exists in the vicinity of the equiatomic
composition. In this system the magnetic Curie temperature
(TCb = 1258 K at xbCo = 0.465) is higher than the
compositional orderdisorder transition temperatures (T0b =
1003 K at xbCo = 0.49) [13].
The Co0.5 Fe0.5 B2 lattice is one of the simplest
structures encountered among the ordered alloys [1]. In
the disordered state, atoms of one type (Fe or Co) are
distributed with equal probability over all sites of the bcc
lattice. In a completely ordered state for stoichiometric
composition all the sites at the corners of the cubic unit
cells are occupied by Fe atoms and the sites at the
centres of the cells are occupied by Co atoms. The LRO
parameter in such an alloy decreases continuously with
increasing temperature and becomes zero at the critical
temperature, i.e. a second-order phase transition occurs at
this temperature. One can understand that in the h100i
direction each ideal bulk (100) plane contains only one
type of atom leaving either all Fe atoms at the surface or all
Co atoms. Hence, there are two equivalent types of bulk

Surface effects

termination: Fe termination (. . . ABAB . . . stacking) or Co


termination (. . . BABA . . . stacking), while the (110) and
(111) planes contain both Fe and Co atoms arranged in a
two-atom unit cell (mixed FeCo termination).
So far the surface structure and composition of the
ordering CoFe alloys have not been studied in so much
detail as observed above binary alloys, although several
surface studies of this system were performed under UHV
conditions in order to determine the surface composition
[4750]. Unfortunately in all experimental works only
the polycrystalline samples were used. Nevertheless we
consider those data which were measured after annealing
at relatively high temperature below and above T0b in
order to approach the equilibrium state. The first result
in this connection was reported by Moran-Lopez and Wise
[47] using AES. After annealing at 723 K (below T0b )
the polycrystalline Co0.5 Fe0.5 sample exhibited a surface
composition equivalent to 755 at.% Fe and 255 at.% Co,
i.e. considerable enrichment with Fe (estimated escape
or about 3 to 4 atomic layers).
depth was about 10 A,
Later, Mrozek et al [51] using quantitative formalism
converted AES spectra into Fe surface concentration of the
polycrystalline Co70 Fe30 , Co50 Fe50 , Co40 Fe60 , and Co30 Fe70
alloy. No change was found in Fe surface composition
at the surface of the Co50 Fe50 alloy in the temperature
range of 683 to 1123 K. On the other hand, moderate Fe
surface enrichment after annealing at 973 and 1123 K was
established for other alloys. Hence these results excluded
the possibility of a considerable Fe or Co segregation effect
at the surface region of the CoFe alloys.
3.2. Ni3 Fe
The FeNi alloy system is of metallurgical interest because
it is the basis of an important class of engineering materials.
Particularly, the investigation of this alloy primarily has
aroused considerable interest due to its special magnetic
properties. For example, the Fe35 at.% Ni alloy (Invar
alloy) has an almost zero thermal expansion coefficient over
several hundred degrees above room temperature; the alloy
with 78 at.% Ni (Permalloy) is a soft magnetic material with
high magnetic permeability. Depending on the composition
and heat treatment conditions, these alloys can have various
structures and, therefore, possess different properties.
In the phase equilibrium state Fe and Ni form a
continuous series of solid solutions [13] and the alloys
important for ferromagnetism lie in the composition range
from 30 to 90 at.% Ni including an (bcc) + (fcc) or
pure fcc region. The orderdisorder transition covers
a wide range of composition, about 5080 at.% Ni. In a
completely ordered alloy the stoichiometric composition is
consistent with Ni3 Fe phase which has fcc Cu3 Au (L12 )
type structure. For this composition the first-order phase
transition occurs at the critical temperature T0b = 776 K.
The phase diagram also shows the composition-dependent
magnetic transition with the maximum Curie temperature
TCb = 885 K at 70 at.% Ni, i.e. TCb > T0b in the FeNi
system.
Several studies on the surface of FeNi alloys have been
preformed [5254]. However, we consider only results

Figure 19. The energy dependence of the differentiated


(00) reflex LEED intensity (dI00 /dE0 ) for Ni3 Fe(111);
1350 K, 2550 K, 3800 K [55].

obtained from single crystals in UHV, in particular the


ordered alloy with Ni3 Fe composition.
3.2.1. Surface structure and layer order parameter.
The first study of the structure and composition at the (100),
(110) and (111) faces of ordered Ni3 Fe alloy was performed
by Vasiliev and Gorodetsky [53] using qualitative LEED
and AES. The surfaces of the single-crystal specimens
were cleaned by the conventional procedure of Ar ion
bombardment and annealing. After a long-time annealing
at 700 K the cleaned surfaces oriented in the [100], [110]
and [111] directions yielded only a (1 1) LEED pattern,
typical for an unreconstructed surface, in the temperature
range from room temperature up to 1025 K (>T0b ). It
must be emphasized that in the case of the Ni3 Fe alloy
it is principle difficult to observe ordered superstructure by
conventional LEED, because the Ni and Fe atoms have
closely related scattering atomic factors in contrast to the
AuCu system. In order to resolve this problem Vasiliev
and Gorodetsky [55] have pioneered applying the procedure
of differentiation of the LEED I (V ) spectra (figure 19).
One can see in figure 19 the clear Bragg and multiplereflection peaks in the differentiated I (V ) spectra. It
was supposed that the interrelation between temperature
m
dependence of the multiple-reflection intensities Ihk
and the
change of the LRO parameter was
m
(s )2 e2M .
Ihk

(4)

For example, figure 20 shows temperature behaviour


for several electron energies E0 under the assumption that
s = b = 1. Let us discuss these results later with
consideration for the temperature dependence of the surface
composition data.
3.2.2. Temperature-dependent layer composition. It is
known that the ordered Ni3 Fe alloy has an fcc lattice with
3049

M A Vasiliev

Figure 20. The temperature dependence of the LRO


parameter for Ni3 Fe(111), 14 surface layers.

the Fe atoms preferentially at the cube unit cell corners and


Ni atoms at the face centres. Consequently, the ideal Fe
atom fraction in (100) or (110) planes of the ordered bulk
crystal is alternately 0.5 and 0, while each (111) plane is
of stoichiometric composition (0.75 Ni + 0.25 Fe).
The first AES observations by Vasiliev and Gorodetsky
[53] of the low-index faces of the ordered Ni3 Fe single
crystal showed only the average composition of the
various surface layers. Nevertheless, this study has made
it possible to reveal some unconventional behaviour of
the temperature-dependent surface segregation (figure 21)
from room temperature (ordered phase) to 1200 K
(disordered phase). One can see from this figure that
no practically changes in Fe surface concentration took
place in the temperature range of 300 to 600 K. The
nearly identical atomic radius of the metal atoms and
small diffusion coefficient at low temperatures in the
FeNi alloys point to relatively small contributions to
disordering processes and consequently to preferential
surface segregation. Subsequent temperature increase leads
to the strong enrichment of Fe up to a temperature
close to the bulk orderdisorder transition temperature
T0b = 776 K. Above this temperature the Fe content
is somewhat decreased in the temperature range of the
magnetic transition but then the surface composition does
not vary up to 1025 K. Despite the general nonmonotonic
behaviour of the Fe content as a function of T , for all
three observed faces, there is a strong orientation effect.
For example, in the highest temperature region studied,
the surface average concentrations of Fe atoms are 35,
48 and 60 at.% for the (100), (110) and (111) faces,
respectively. One remarkable result is the observation
of the surface segregation maxima for all faces at the
same temperature close to T0b = 776 K. It was then
reasonable to suppose that surface segregation is related
to the bulk orderdisorder transition, and the temperature
at the onset of segregation, 645 K, corresponds to the
beginning of the composition disordering in the surface
region. In this case one would expect a damped oscillatory
behaviour from the concentration profile in a direction
perpendicular to the surface as expected for an ordering
alloy. In order to investigate the depth dependence
of Fe concentration Vasiliev and Gorodetsky [56] have
developed a novel approach of a non-destructive layer-bylayer composition analysis. The principle of the approach
is the experimental measurements of the relation between
3050

Figure 21. Temperature dependences of the Fe surface


concentration for Ni3 Fe(100), (110) and (111) faces [53].

surface composition, determined by means of the ionization


spectroscopy (IS), and the energy of primary electrons,
with subsequent recovery from these data of layer-bylayer information by mathematical processing methods.
As an example, figure 22 shows the layer dependence
of the Fe distribution at different annealing temperatures
for Ni3 Fe(111) alloy. One can see from this figure the
strongly damped oscillatory character of the concentration
profile in the range of disordering temperatures. It is
apparent that the amplitude of oscillations decreases at
T > T0b . Thereafter the composition-corrected LRO
parameter was determined as a function of both temperature
and depth (figure 23). We note the following main results
obtained from this study. (i) In contrast to the abrupt
change at critical temperature of the bulk LRO parameter
b , the corresponding parameter for the surface layers is
(i = 1, 2, 3, . . . layer) decreases smoothly with increasing
temperature. This means a continuous behaviour of the
surface orderdisorder transition like a second-order type
one. (ii) The surface disordering begins earlier that in the
bulk: for the first surface layer the critical temperature is
nearly 75 K below the bulk one, while for deeper layers this
temperature coincides with the bulk one. (iii) From layer
to layer the kind of orderdisorder transition changes from
second order to first order. (iv) There is strong competition
between segregation and ordering processes.
3.2.3. Ferromagnetic anomalies. It is well known that
magnetic phase transitions can have a pronounced effect on
physical, chemical and other bulk properties of magnetic
materials [57]. There is also evidence for such an effect
at the surface of pure ferromagnetic metals. The best
known example in this respect is the nickel surface which
exhibits several ferromagnetic anomalies in the vicinity of
the Curie point. In particular, magnetic scattering and
polarization of electrons from the Ni(100) face near the
critical region [58, 59], new morphological transition at
the Curie temperature [60], reconstruction of stepped Ni
surfaces [61], reversible step rearrangement and segregation
on Ni surface at T0b [62], and emission of secondary
particles during ion sputtering of ferromagnetic metals in
the phase transition region [194, 195] was observed.

Surface effects

Figure 22. The layer dependence of the Fe concentration


at different temperatures for Ni3 Fe(111) [56].

Figure 23. The layer dependence of the LRO parameters


at different temperatures for Ni3 Fe(111): theory [145],
LEED experiment [146].

In our opinion the first surface ferromagnetic anomalies


in alloys were reported by Vasiliev and Gorodetsky [63] and
Mamaev et al [64] for the clean surface of an Ni3 Fe single
crystal. The first effect was observed [63] in the course of
the continuous registration of the temperature dependence
of Bragg maximum intensity Ihk (T ). This experiment
was made possible by an improved method of LEED
reflex modulation and automatic adjustment of electron
beam energy on the maximum of the reflex intensity [55].
Thus it was revealed that for both pure Ni and Ni3 Fe
the temperature dependence of the Bragg reflex I00 (T )
exhibited nonmonotonic behaviour in the region of the
Curie temperature TCb . Figure 24(a), (b) gives an example
of such dependence for Ni(100) and Ni3 Fe(100) surfaces
respectively. The same behaviour of I00 (T ) was also
observed for (100) and (111) faces. It is interesting that
the position of the observed intensity peak near TCb did
not depend on E0 , while its magnitude diminished with

increasing E0 , so that for E0 > 200 eV it could not


be recorded. Consequently, only the outermost surface
layers contributed to this effect, which may be attributed to
competition between the critical scattering of electrons on
magnetic moment fluctuations near TCb and the scattering of
electrons on thermal lattice vibrations according to the well
known DebyeWaller function. In addition it was found
that the energy of the LEED reflex maximum was altered
near the TCb temperature region. An example of such a
temperature-dependent energy shifting 1E0 (T ) is given in
figure 25 for the Ni(100) crystal. The value of 1E0 both for
Ni(100) and also for Ni3 Fe(100) is dependent substantially
on the E0 value. At E0 25 eV the shifting 1E0 amounts
to 3 eV and decreases with E0 rise. It is interesting that
the sign of shifting is different for Ni and Ni3 Fe. So with
temperature increasing the maximum energy for Ni became
lower while for Ni3 Fe it increased during the run across TCb .
There was also an orientation dependence of the 1E0 . Thus
for Ni and Ni3 Fe at the same E0 the shifting 1E0 was two
times larger for the (100) face than for (111). It is apparent
that from the quantity 1E0 in the case of beam spicular
reflection (00), it is possible to determine the change of the
interplane spacing at normal to the surface: 1d/d = (1/2)
1E0 /E0 . For example, the relative change of this spacing
at E0 = 30 eV (surface region) was 5 102 , while at
E0 = 200 eV 1d/d 2 105 which is close to the
values typical for magnetostriction processes in the bulk.
Finally it was suggested that the shifting effect 1E0 in the
TCb region is a result of competition between ordinary lattice
thermal expansion and magnetostriction (thermostriction)
effects which as well known are different in Ni and Ni3 Fe
crystals [3].
Mamaev et al [64] first studied the magnetic properties
of the clean ordered Ni3 Fe(111) alloy surface using
SPLEED. Figure 26 shows the temperature dependence of
the scattering asymmetry of the spin-polarized electrons
ASE (T ). The change of the asymmetry sign in the vicinity
of TCb , i.e. the change of the magnetic coupling type, is
clearly seen. The Curie temperature of the (111) surface
for the Ni3 Fe alloy TCs alloy appears to be 1050 K, which is
much higher than the TCb = 850 K. Thus the present work
first revealed the antiferromagnetic coupling between the
Fe-rich surface layers and the underlying bulk of Ni3 Fe.
Recall that the strong Fe segregation at high temperatures
for this ordered alloy was mentioned above.
4. Intermetallic ordered compounds: surface
layer structure and composition
4.1. Fe3 Al(110)
AlFe alloys are potentially useful for high-temperature
industrial applications. The phase diagram of the AlFe
system shows the existence of two ordered phases based on
the ideal composition Fe3 Al (2 ) and Fe0.5 Al0.5 (2 ) [13].
The Fe3 Al superlattice is of the bcc BiF3 (DO3 ) type, with
twice that of the random
a lattice constant (a0 = 5.84 A)
phase. In this structure the atoms of Al avoid adjacent
sites and so occupy the centres of the alternate small cubes
(each 1/8 of the true unit cell) that show the structure of
3051

M A Vasiliev

(a)
Figure 26. The temperature dependence of the scattering
asymmetry of spinpolarized electrons for Ni3 Fe(111) [64].

(b)
Figure 24. LEED intensity of the (00) reflex I00 and the
intensity logarithm ln I00 versus temperature corresponding
to (a) Ni(100) and (b) Ni3 Fe(100) [63].

Figure 25. The temperature dependence of the energy


shifting for the LEED (00) reflex [63].

pure Fe. The Fe3 Al crystal exhibits a continuous order


disorder phase transition (DO3
B2) with a concentrationdependent (in the range 2434 at.% Al) critical temperature
TCb of 800 K. The Curie point drops with addition of Al,
and alloys containing more than 35 at.% Al are no longer
ferromagnetic at room temperatures [3, 13].
The first study of the surface properties for such a
3052

complicated ordered compound as Fe3 Al was performed


by Mailander et al [193]. They have measured the nearsurface critical glancing angle x-ray scattering at the (110)
face of an ordered Fe3 Al single crystal. After Ar ion
sputtering and annealing at T < 800 K the surface showed a
sharp (11) LEED pattern. Using then the effect of total xray reflection within the temperature range between 300 and
800 K the surface parallel exponent (k = 1.52 0.04) was
determined. The observed exponent agrees reasonably with
the predicted value k = 1.48 for the so-called ordinary
transition [65, 66]. It was also confirmed that the lateral
correlations near the surface decay much faster than at
As another interesting
a depth of the order of 100 A.

result, an experiment with a small escape depth of 30 A


showed the existence of a superlattice Bragg peak up to
temperatures of 16 K above T0b . It was assumed that
existence of the surface long-range DO3 ordering even for
the disordered B2 phase could be connected to surface
segregation which can cause changes in the transition
temperature according to the AlFe phase diagram.
The first surface segregation experiments on Fe3 Al(110)
near the orderdisorder transition temperature were
performed by Voges et al [67] using LEIS. The surface
of the Fe0.71 Al0.29 (110) single crystal was cleaned by Ar
ion sputtering, followed by a final annealing at 600 K. The
LEIS measurements at room temperature demonstrated a
slight Al enrichment in the surface; however the interatomic
distances were compatible with the ordered bulk lattice
parameters. Quantitative analyses have shown, that Al
concentration for the first layer in the [111] direction
x1Al = 0.38 and in the [110] direction x1Al = 0.43. The
increasing equilibrium Al content for elevated temperatures
was also clearly shown by the ISS data (table 1). But in
so doing no evidence was found for a change in these Al
distributions by crossing the bulk transition temperature.
The results depicted in the table exhibit a slight Al depletion
in the second atomic layer, illustrating the oscillatory
concentration profile.

Surface effects
Table 1. Al concentrations in the first and second
outermost layers, estimated from ISS [67].
Temperature (K)

x1Al

x2Al

700
900
1000
1050

0.96 0.01
0.97 0.01
0.97 0.01
0.98 0.01

0.0 0.1
0.1 0.1
0.3 0.1
0.2 0.1

4.2. Ni3 Al
The AlNi system has received recently considerable
attention because of its technological importance. Some
ordered intermetallic compounds, especially Ni3 Al and
NiAl, yield strength increases with increasing temperature
and have excellent corrosion resistance at high temperature
[68]. However the intrinsic brittleness of this compound
in its polycrystalline form is a major drawback to its
use as a construction material. It was suggested that the
local compositionally disordered regions close to the grain
boundaries are liable to increase the tendency for cracking
to occur, since the number of possible dislocation reactions
at the interface is larger for the disordered state. It may be
that this local disorder is due to surface segregation effects.
It is one of the main reasons for the current experimental
interest in the surface structure of such ordered compounds
in single-crystal form. Since AlNi alloys contain both
simple and transition metals, they also provide a challenge
of two intermetallic compounds of the AlNi system have
been examined in detail: Ni3 Al (fcc, Cu3 Al L12 type) and
NiAl (bcc, CsCl B2 type), with different low-index faces
[13]. Let us note that both compounds are paramagnetic [3].
4.2.1. (100) face. There are two possible bulk terminations: in the {100} termination (denoted by A) the top
atomic layer has 50 at.% Ni50 at.% Al (mixed layer);
in the other termination the {200} planes (denoted by B)
contain only 100 at.% Ni atoms. In order to determine the
type of termination Sondericker et al [69, 70] performed the
first quantitative LEED intensity analysis of the Ni3 Al(100)
face. In the first place, it was indicated by AES that a
cleaned and annealed surface was stoichometrically stable
between room temperature and 1073 K, i.e. a surface segregation does not take place here. In the second place, the
LEED-intensity-calculated data on the basis of the dynamical multiple-scattering program and R-factor analysis confirmed the mixed-layer termination of a clean Ni3 Al(100)
surface with a small contraction of 1d12 = 2.8% (db =
and a slight buckling of the top layer (the Al atoms
1.78 A)
outwards from the Ni atoms). The
being 0.02 0.03 A
second interlayer spacing appeared to be bulklike. Firstprinciples calculations of the cohesive energies of slabs terminating in the two types of layer also indicated that the
mixed-layer termination is more stable (2.108 Ryd compared to 2.036 Ryd for the B termination) [70].
4.2.2. (110) face. As in the case of Ni3 Al(100) there
are also two possible {110} terminations: (i) unit mesh
with one Al and one Ni atoms in the first layer and two

Ni atoms in the second layer (mixed-layer termination);


(ii) unit mesh with two Ni atom in the first, and one Al
and one Ni atom in the second layer (Ni-layer termination
indicated). This question of true termination was resolved
by Sondericker et al [71] in their quantitative LEED
analysis. The clean and annealed (110) face up to 973 K
revealed that the topmost layer contains 50% Ni50% Al
and the second layer has 100% Ni. On the basis of the
LEED intensity calculations the geometrical parameters
were also determined. Particularly in the first surface
layer, the Ni and Al subplanes were slightly separated
(1.2% of the bulk
from one another by 0.02 0.03 A
the Al atoms moved outwards from the
db = 1.259 A),
bulk. The contraction of the distance between the Ni
(or 12%
subplane and the second layer was 0.15 0.03 A
of the bulk value, while the corresponding value for the
Al subplane was 10.7%). The second interlayer distance
(3% of the bulk
was expanded by 1d23 = +0.04 0.03 A
value). Buckling of the second plane and deeper relaxations
have not been tested.
4.2.3. (111) face. In this case there is only {111}
termination because all {111} planes are equal to one
another, with the stoichiometric composition of three Ni and
one Al atom per unit cell. The structure of the Ni3 Al(111)
face was determined by Sondericker et al [72] by means of
a quantitative LEED intensity analysis assisted by AES. It
was confirmed that bulklike surface composition was stable
between room temperature and 1023 K after the surface
cleaning procedure. The following structure parameters
were determined: a small but detectable buckling of the
above
first atomic layer with the Al atoms 0.06 0.03 A
the plane of the Ni atoms, which is in turn very slightly
toward the second atomic
shifted inward (0.01 0.03 A)
layer; second and deeper interlayer spacings were expected
Hence the
to be equal to the bulk value (db = 2.055 A).
termination of the Ni3 Al(111) single crystal is essentially
bulklike but with a small buckling of the topmost layer.
4.3. Ni0.5 Al0.5
4.3.1. (100) face. Since the Ni0.5 Al0.5 compound has
the CsCl type ordered structure all bulk {100} planes have
an . . .ABAB. . . stacking sequence with layers of all Ni
atoms and layers of all Al atoms alternately interspaced,
and both layers have a square unit cell. Hence there is
a question concerning bulk termination for a real crystal
of Ni0.5 Al0.5 (100): an Al layer or Ni layer. This surface
was first examined by Davis and Noonan [74] using the
quantitative LEED analysis. The calculated I (V ) spectra
were compared with the experimental spectra by use of the
reliability R-factor. The lowest R-factor was found for the
Al-terminated model. It was also determined that surface
relaxation values are 1d12 = 8.5% and 1d23 = +4.0%

(db = 1.44 A).


LEIS and LEED studies on the same surface by Mullins
and Overbury [73] have confirmed the Al-termination
model, although also a surface reconstruction was observed.
Depending upon the annealing regimes twotypes
of LEED
pattern were observed in this study: (i) a c( 23 2)R45
3053

M A Vasiliev

LEED pattern was formed by annealing the Ar-sputtered


surface at 750 < T < 950 K (this type of pattern is generally indicative of a 2:1 surface composition with like atom
in rows along the [011] or [011] direction); (ii) annealing the crystal to T > 1000 K irreversibly produced a
(1 1) LEED pattern. The true first-layer composition for
every kind of surface structure was determined from the Li+
scattering
intensity. It appeared that for the reconstructed

c( 2 3 2)R45 phase the first-layer atom fraction of Ni


was 0.22 0.07 and Al was 0.78 0.07. It is not pure
Al termination as was calculated in the work [74]. Finally
and Overbury [73] have calculated that the
Mullins
c( 2 3 2)R45 structure is attributable to a compositional order of mixed Al and Ni adsorbed in fourfold sites
above a second layer composed almost entirely of Ni atoms;
and the (11) structure is associated with a random mixture
of Ni, Al and a small fraction of vacancies above a second
layer composed of Ni. It was also pointed out that a general trend toward decreasing Al surface composition in the
case of Ni0.5 Al0.5 (100) is a result of the competition of two
processes: the strong ordering because of the very exothermic heat of mixing in Ni0.5 Al0.5 (14 kcal mol1 ) which
favours each component being surrounded by the other
species, i.e. the bulk structure; and tendency to Al segregation due to its heat of vaporization being lower than that of
Ni (67.9 kcal mol1 versus 90.5 kcal mol1 ) and its atomic
versus 1.62 A)
as expected from
size being larger (1.82 A
quasi-thermodynamic models of surface segregation [75].
4.3.2. (110) face. Since the Ni0.5 Al0.5 compound has the
CsCl(B2) structure, all ideal bulk {100} planes contain half
Ni atoms, which are coplanar and ordered with a rectangular
unit cell. The surface structure of the actual Ni0.5 Al0.5 (110)
single crystal was investigated by different experimental
techniques, such as LEED, MEIS, LEIS and AES.
The first quantitative LEED study on the Ni0.5 Al0.5 (110)
face by Davis and Noonan [74, 76] showed the bulk
termination, mixed 5050 NiAl composition after
annealing at 1183 K. The experimental I (V ) spectra
compared with dynamical calculated spectra (using the Rfactor) have first demonstrated a surprising result: the
outermost (110) layer possesses a relatively large rippled
relaxation.
So Ni atoms contract into the bulk by
1d12 (Ni) = 6.0% (relative to bulk interlayer spacing
and Al atoms expand to the vacuum region by
2.04 A)
1d12 (Al) = +4.6%. These values 1d12 correspond to a
rippled first composite layer, where the Al sites are above
In the next work [74] Davis and
the Ni sites by 0.22 A.
Noonan considered the rippling in the second composite
layer, and proposed a final model (figure 27): Al atoms of
the first layer relaxed outward by 5.2% and the Ni atoms
relaxed inward by 4.6%, i.e. a total rippling of 9.8% or
in the second layer Al atoms relaxed outward
0.20 A;
by 1d23 (Al) = 2% and the Ni atoms relaxed outward by

1d23 (Ni) = 1%, i.e. a total rippling of 1% or 0.02 A.


Ni0.5 Al0.5 (110) has been also investigated by use of
an MEIS technique with a channelling and blocking by
Yalisove and Graham [77]. They result are in excellent
agreement with the LEED analysis [74, 76].
3054

Figure 27. Side view of a ball model for the rippled


Ni0.5 Al0.5 (110) surface [74].

The composition of the ordered Ni0.5 Al0.5 (110) surface


was also determined by Mullins and Overbury [73] from
LEIS with Li ion scattering. It was confirmed that a (100)
surface well annealed at 1000 K was composed of equal
amounts of Ni and Al atoms. Hence surface composition
in the case of the (110) surface indicated that the high heat
of mixing dominates any tendency toward segregation.
Wuttig [78] has investigated the dispersion of surface
phonons for the (110) surface of the ordered Ni0.5 Al0.5
compound and confirmed the rippled structural model based
on LEED study.
4.3.3. (111) face. The {111} bulk planes of the ideal
ordered Ni0.5 Al0.5 crystal expose a very open structure with
alternating Ni and Al layers in an . . .ABAB. . . stacking

sequence and a small interlayer spacing of only 0.83 A.


Consequently, the termination of the bulk structure of the
real crystal in this case may result either in an Ni- or an
Al-terminated surface, with the atoms laterally arranged
in hexagons. Indeed, from the first observed (1 1)
LEED pattern of the Ni0.5 Al0.5 (111) face by Noonan and
Davis [79] it was concluded that the surface has the
same lateral unit mesh as compared with the bulk plane.
The experimental I (V ) spectra (at T < 1273 K) were
compared with spectra resulting from dynamical LEED
intensity calculations, and the conclusion was made that
there is a 50%50% mixture of both Al-terminated and
Ni-terminated compositionally ordered domains. Such a
model implicitly implied the existence of single atomic
steps separating the Ni and Al areas by going up and
down in the . . .ABAB. . . stacking sequence. Additionally
different multilayer relaxations in the two domains were
determined: 1d12 (Al) = 5%, 1d23 (Al) = +5%,
1d12 (Ni) = 50%, and 1d23 (Ni) = +15%. Noonan
and Davis [79] have given a possible explanation of the
observed Ni0.5 Al0.5 (111) surface based on the assumption
that the surface free energies of the two possible bulk
terminations are comparable.
However, Niehus et al [80] have proposed another
explanation of the Ni0.5 Al0.5 (111) surface using LEIS and
STM. The surface when annealed below 1300 K was found
to consist of large flat terraces separated by double atomic

Surface effects

steps and having only one atom species (Ni). These


terraces were covered with small domains (an average
of the other alloy component (Al) which
size of 25 A)
might be segregated over the top layer under the influence
of observed oxygen (about 10% of a monolayer). On
this surface Al-terminated terraces were not found. Hightemperature annealing at 1400 K resulted in an oxygen-free
surface and there was only Ni termination of the first layer
followed by a second layer consisting of Al atoms.
4.4. Pt3 Ti(100) and (111)
Alloys of Pt and Ti atoms are characterized by a
highly negative enthalpy of formation and strong bonding
interaction in the d orbitals of each metal. PtTi alloys
are known as a compound-forming system, particularly,
Pt8 Ti, Pt3 Ti, Pt3 Ti2 , PtTi and Ti3 Pt ordered compounds
exist according to the bulk phase diagram [13]. The
strong intermetallic bonding interaction is expected to cause
dramatic changes in the catalytic properties of Pt and Ti.
The ordered surface Pt3 Ti compound is one of the stable
systems that forms among transition metals. The ordered
surface structure for this compound was the object of
several works [8184].
The bulk single-crystal structure Pt3 Ti has the Cu3 Au
L12 type structure with Ti atom substructure for Pt at
the corners of the fcc cubic unit cell, producing an
. . .ABAB. . . sequence of planes normal to [100]. These
{100} planes alternately have 50 at.% Ti and 100 at.% Pt
composition. Thus there are as usually two inequivalent
regular terminations possible. On the other hand, an ideal
(111) surface of this crystal should have a layer stacking
designated as . . .ABCABC. . . and all layers should be of
identical stoichiometric composition, i.e. three Pt atoms and
one Ti atom per unit cell.
Early work on the low (100) and (111) faces of
Pt3 Ti was performed by Bardi and Ross [81] using
qualitative LEED and AES. LEED patterns for the
clean annealed (111) and (100) faces were designated as
p(2 2) and c(2 2) superlattices, respectively. The halforder (superlattice) spots were observable to the highest
temperature up to 1200 K. After examining the superlattice
spots over a wide range of beam energies and Ti/Pt
Auger peak ratios Bardi and Ross [81] concluded that the
Pt3 Ti(100) face is a mixed bulklike (100) layer, and the
(111) face is again the bulk termination. However, later
Paul et al [82], using angle-resolved x-ray photoemission
and LEIS, indicated that the clean surface of Pt3 Ti(111)
has a reconstructed structure where the first atomic layer
is a quasi-pure Pt, and the stable termination of the
(100) surface is the Pt-rich one. These results have been
confirmed by a complete LEED analysis for the (100)
face by Atrei et al [83] and for the (111) face by Chen
et al [84]. In the latter more detailed study the full
quantitative analysis of LEED I (V ) spectra was performed
over several structure models. It was established as result
that one pure Pt layer atop the bulk lattice gave the best
agreement with experiment on the base measured R-factors,
i.e. the outermost layer was pure Pt and the other layers
had the bulk Pt/Ti ratio 3:1. The small contraction of

0.03 A
for the first interlayer spacing (d12 =
0.02 A
and 0.04 A0.03

for the second interlayer


2.230.03 A)
A
corresponding to contraction
spacing (d23 = 2.210.03 A),
of 0.9% and 1.8% of the bulk value, respectively, were
determined. It was also found that there is a small buckling
0.05 A
and larger buckling in the
in the top layer 0.04 A
0.04 A.

second layer 0.15 A


4.5. Pt3 Sn
Surface properties of the highly exothermic bulk PtSn
alloys have been extensively studied in view of the interest
as both heterogeneous catalysis for hydrocarbon conversion
and as electrocatalysts for the direct electro-oxidation of
methanol in fuel cells. Moreover, this system is interesting
for the study of surface segregation because it is composed
of metals of considerably different heat of sublimation
and surface energy [85]. The phase diagram of the Pt
Sn alloys shows that the solid solubility of Pt in Sn is
negligibly small, and assumes the existence a number of
stable compounds, particularly Pt3 Sn, PtSn, Pt2 Sn3 and
Sn2 Pt [13].
Several recent studies of the surface structure of a
Pt3 Sn single crystal were performed by surface sensitive
techniques [83, 86, 87]. The highly ordered exothermic
compound Pt3 Sn has the Cu3 Au (L12 ) structure with Sn
atoms on the corners of the face-centred cubic unit cell and
Pt atoms on the centre of the faces. The ideal faces of this
crystal must have bulk termination surfaces like the other
members of the L12 family, for example, such as Cu3 Au,
Ni3 Al and Pt3 Ti. Because of the surface Sn enrichment
observed in polycrystalline Pt3 Sn [88] and the relatively
large difference in the thermodynamic parameters between
Sn and Pt, the proposal that Pt3 Sn will behave like Cu3 Au
has come into question.
4.5.1. (100) face. In the ideal (100) face of the Pt3 Sn
single crystal there are equal numbers of the two types
of atom (50% Sn50% Pt) and each has eight nearest
neighbours. In the (200) plane there are only Pt atoms,
each with eight nearest neighbours. Consequently, in
the [100] direction surface planes alternate between a
mixed layer and pure Pt one, resulting in two inequivalent
regular terminations of the bulk crystal. The preferential
termination in the ordered Pt3 Sn(100) compound was
first determined by Haner and Ross [86]. After Ar
ion bombardment and annealing at 1032 K the well
ordered c(2 2) LEED pattern was observed and LEIS
spectra clearly indicated a surface composition of at
least 50 at.% Sn. Those results are consistent with the
termination of the compositionally mixed (100)-c(2 2)
structure. In subsequent experiments Atrei et al [83] using
a quantitative LEED intensity analysis provided support for
this conclusion. In addition to LEED pattern observations
they have performed full dynamical calculations of the
intensity of the diffracted LEED beams, considering the
variations in the first two interplanar distances (d12 and d23 )
and the buckling of Sn atoms in the first mixed surface layer
with c(2 2) structure. Assuming the outlined termination
with 50% of Sn, the best agreement between experiments
3055

M A Vasiliev

5. Random solid solution


5.1. Ordering solid solution

Figure 28. Top view of the surface structure of Pt3 Sn(111):


(a) bulk termination; (b) surface reconstruction [87].

and calculations was found for buckling of Sn atoms of


a value of d12 = 1.89 0.05 A
and a value
0.22 0.08 A
(bulk value).
of d23 = 2.00 0.05 A

4.5.2. (111) face. The ideal (111) face contains three


Pt atoms for every Sn atom, and has a composition
identical with the average of the bulk. Each atom here
has nine nearest neighbours. Hence the {111} planes in
the Pt3 Sn crystal are all equivalent and contain the bulk
composition. A structural model for the (111) surface of
the ordered Pt3 Sn compound corresponding to the ideal
bulk truncation (figure 28(a)) was confirmed by Haner and
Ross [86] and Atrei et al [83] using quantitative LEED
and LEIS analysis. In these works the LEED patterns have
been indexed as p(2 2). Full dynamical calculations
of the LEED beam intensity have shown stoichiometric
composition (25 at.% Sn) of the outermost layer, Sn atom
d12 = 2.20 0.05 A
and
buckling of 0.21 0.08 A,
(db = 2.31 A).

d23 = 2.38 0.08 A


Recently Atrei et al [87] have demonstrated that the
(111) surface of the Pt3 Sn ordered compound may show
a reconstruction in some conditions of preparation, i.e.
a non-bulk termination. So after the clean surface with
(2 2) structure was sputtered by 3 keV Ar ions at
600
K the new
LEED pattern was observed corresponding
to ( 3 3)R30 . Based on the quantitative LEED
crystallographic analysis, Atrei et al [87] proposed the
structural model of the reconstructed Pt3 Sn(111) surface
(figure 28(b)). This model consists of a topmost plane
where Sn atoms substitute

Pt atoms in an orderly way


according to the ( 3 3)R30 hexagonal symmetry,
while the underlying layers are assumed to contain no Sn.
The optimization of the structure parameters led to a slight
of the Sn atoms in the overlayer.
upward buckling (0.2 A)
The Sn fraction in the topmost layer for both (2 2) bulk
termination and reconstructed model is, respectively, 1/3
and 1/4 of a monolayer.

It was also concluded that the


main factor of the ( 3 3)R30 phase stabilization was
the highly negative enthalpy of Pt3 Sn formation [85], that
may be attributed to the enhanced PtSn nearest-neighbour
bonding. It takes place when the subsurface layers are
depleted in Sn. In this case Sn atoms are surrounded by Pt
atoms only, and in contrast to the (2 2) bulk termination
for the reconstructed Pt3 Sn(111) face the SnSn nearestneighbour pairs between atoms of the first and the second
surface layers do not exist.
3056

5.1.1.
Cu85 Pd15 (110). Alloys of Cu and Pd are
interesting first of all in the field of catalysis and have been
the object of a number of surface studies [86, 8995]. Cu
and Pd atoms form a continuous series of solid solutions
[13], while the enthalpy of mixing in the formation of the
alloys is only a moderately negative. Since the atomic
radii of Cu and Pd differ by about 7%, the CuPd system
is an appropriate alloy for investigating the influence of
atomic strain on the surface segregation effects. What is
more, those alloys make up the compositional ordered bulk
phases in certain bulk composition ranges [13]. The ordered
superstructure based on Cu3 Pd (T0b 773 K) is that of
the fcc Cu3 Au (L12 ) type within the composition range of
about 1025 at.% Pd and the orderdisorder transition in
the region 3050 at.% Pd is associated with formation of
the CuPd ordered superstructure (T0b 873 K) of the bcc
CsCl type. Evidence for surface ordered phases in this
system was first found by qualitative LEED measurements
for CuPd(100) and CuPd(110) alloys which were formed
by depositing Pd atoms on the clean surface of the Cu
crystal followed by annealing [89]. In this case Fujinaga
found the c(22) ordered structure on the (100) surface and
the (21) structure on the (110) surface in agreement with a
bulk termination structure of the Cu3 Pd ( 0 ) ordered phase.
Anderson et al [90] and Pope et al [91] first described
the kinetics of the domain growth during formation of the
Cu(100)c(2 2)Pd surface alloy. Holmes et al [95] using
LEED and AES have extensively analysed the equilibrated
clean surface of a single-crystal Cu85 Pd15 (110) alloy, and
first reported the geometric structure and composition of
this face. It is well understood that a cut along the
(110) plane will give alternate layers of 50:50 CuPd and
100 at.% Cu for the ideal Cu3 Pd phase. However, here
two possible surface terminations may exist: (i) a top pure
Cu layer, and (ii) a topmost mixed CuPd layer. Holmes
et al [95] have found that after repeat cycles of Ar ion
bombardment and annealing at 650 K the LEED pattern
corresponded to a well ordered p(2 1) superstructure.
Based on the observation of CO adsorption it was suggested
that pure Cu-terminated the first surface layer with a
mixed second ordered CuPd layer was the most likely
surface structure (figure 29). Subsequent results presented
by Lindroos et al [96] using quantitative LEED, and by
Newton et al [93] using LEED, XPS and LEIS have
confirmed the conclusion of Holmes et al [95]. However, it
should be stressed that according to the bulk phase diagram
the Cu85 Pd15 alloy does not achieve the ideal ordered Cu3 Pd
structure in contrast to the behaviour in the surface region.
It was the first case of the ordering of an underlayer at the
surface of a random substitutional alloy, associated with the
surface segregation phenomenon.
Rao et al [97] have performed the first extensive study
of UV photoemission (UPS) for the (1 1) disordered
surface of the Cu85 Pd15 alloy. The spectra obtained were
quite different to both Cu(110) and Pd(110) spectra. This is
evidence that the CuPd alloy has been formed with a high
degree of mixing of the Pd and Cu electron states. Both

Surface effects

Figure 29. Schematic diagram of the


Cu85 Pd15 (110)(2 1) surface [93].

ordered and disordered phases of the Cu85 Pd15 (100) face


have been then studied by Newton et al [93] using also
UPS. Figure 30 shows the angle-resolved normal emission
spectra from both the disordered (1 1) and ordered
(2 1) surfaces with HeI radiation (in addition the spectra
from Cu and Pd single crystals are also presented [97]).
Two main changes after annealing to (2 1) structure
were observed: (i) all the features of the photoemission
spectra appear to be better resolved; (ii) the major peak at
2.2 eV binding energy (Cu-derived feature) has increased
in intensity relative to the rest of the spectra. It was
suggested that the first comment is dictated by the observed
ordering of the surface. Rao et al [97] have proposed
to describe the width of both Pd and Cu features in the
(11) HeI spectra in the context of disorder smearing and
damping of the Cu-derived features. The second comment
is consistent with the segregation effect of the Cu atoms
into the topmost layer at the (2 1) selvedge structure
formation. Most recently Cole et al [121] have discussed
the electronic structure of the ordered Cu85 Pd15 (110)
surface in terms of the conclusion drawn from the extensive
experimental and theoretical studies of bulk CuPd alloys
[100, 101, 105, 108, 110, 111, 120]. It was speculated that
a knowledge of the ordering processes in the bulk of
CuPd alloys may yield useful insight into the nature of
surfaces. First-principles electronic structure calculation
(within the framework of the multiple-scattering KKRCPA
approach) for ordered, as well as substitutionally disordered
systems are discussed in a series of papers [98, 99, 102
104, 106, 107, 109, 112, 116, 117]. The above-mentioned
theoretical treatments and experiments have indicated that
the physical and electronic structure (bulk and surface) of
the CuPd alloys can be understood in terms of matrix
element and self-energy effects, local lattice expansions
and charge transfer [114, 121]. The influence of the bulk
chemical order on the surface morphology was recently
demonstrated by Barbier et al [113]. They have shown
a double-stepsingle-step phase transition on the vicinal
surface of Cu83 Pd17 (110). Most recently Gallis et al [122]
have discussed the question of the competition or synergy
between surface segregation and bulk chemical ordering in
the CuPd system. Properties of Cu0.5 Pd0.5 (110) singlecrystal alloy surfaces were recently depicted by LobodaCackovic [118, 119, 123].
5.1.2. PtCo alloys. The PtCo system is one of the
platinum bimetallic alloys studied in terms of both catalytic
and magnetic properties, particularly, in developing highdensity magnetic recording. Pt and Co form a substitutional

Figure 30. Emission HeI valence band spectra


corresponding to Cu85 Pd15 (110)(A) the (1 1) phase and
(B) the (2 1) phase [91]and (C) Pd(110) and
(D) Cu(110) [95].

continuous solid solution in the whole range of composition


and two ordered phases [13]. The PtCo ordered structure
is stable from about 42 to 72 at.% Pt and the maximum
of the orderdisorder transformation temperature is 825 C
at 50 at.% Pt. This phase has the tetragonal structure of
the CuAu (L10 ) type. In the region of 75 at.% Pt there
is a fcc superlattice Pt3 Co of the Cu3 Au (L12 ) type with
T0b 750 C. Both ordered phases are ferromagnetic at
room temperature.
Pt80 Co20 (100). The first qualitative LEED results and
LEIS data by Bardi et al [124, 125] showed that the stable
termination of the (100) surface of the fcc disordered
Pt80 Co20 has a layer of pure platinum, forming a quasihexagonal layer on the (100) alloy surface with parameters
similar to those of the reconstructed Pt(100) surface.
According to LEIS results the second outermost layer
was approximately 40 at.% Co. The observed oscillatory
variation in Pt composition (1006080 at.% Pt) of the first
two layers of the (100) surface is strikingly similar to the
surface-sandwich segregation model proposed earlier for an
ordering system (NiPt, PtTi).
Pt25 Co75 (110). The structure of the Pt25 Co75 (100)
alloy in the substitutionally disordered state was studied
in detail by Bugnard et al [126] using quantitative
LEED analysis.
In the Co-rich region of the Pt
Co phase diagram the hexagonal cobalt phase () is
stable at room temperature until more than 6 at.% Pt is
dissolved, whereupon the structure becomes disordered (up
to 42 at.% Pt) fcc -phase. In this region there is no order
disorder transformation; however the Pt25 Co75 composition
is strongly ferromagnetic with rather high Curie temperature
TC 1120 K. After simple standard Ar ion cleaning and
annealing at 1070 K the Pt25 Co75 (110) crystal exhibits
a sharp (1 1) LEED pattern with no evidence of
compositional ordering on a long-range scale. LEED
I (V ) intensities were calculated using the average T -matrix
approximation for substitutionally disordered alloys. The
goodness of the fit was assessed by five metric distances.
3057

M A Vasiliev

Finally, it was found the stronger Pt concentration oscillates


around the bulk value and damps out over five layers:
the top layer is a pure Co layer (x1P t = 0), the second
layer is strongly Co depleted (x2P t = 90 at.%), the third
layer is strongly Pt depleted (x3P t = 0), the fourth layer is
enriched in Pt and the fifth layer is slightly Pt depleted.
Additionally, larger interlayer spacing oscillations were
found: 1d12 = 16.3% (contracted) and 1d23 = +11.2%

(expanded) compared with db = 1.298 A.


From these data one can notice that the damping of
the interlayer spacing is faster than that of the composition.
Interestingly, the present results demonstrated that on the
three surface region layers of Pt25 Co75 (110) there is quasiperfect stacking of pure metal layers. This is a very
important observation with respect to the surface magnetic
properties of this ferromagnetic system.
Pt80 Co20 (111). In the case of the (111)-oriented
Pt80 Co20 [124], no reconstruction was observed in
qualitative LEED, but only the fcc (111)-(1 1) pattern
expected for a bulk truncation surface structure with a
surface lattice parameter nearly the same as that of the pure
Pt(111) surface; no reflexes attributable to ordering of the
Co atoms in the surface lattice were detected.
Gauthier et al [11] performed the first study of the alloy
Pt80 Co20 (111) surface using quantitative LEED analysis
in the T -matrix approximation. According to the phase
diagram for Pt contents larger than 80 at.% the alloy exists
in the disordered fcc -phase at all temperatures. On
cleaning by standard Ar ion cleaning and annealing in the
range 9601060 K the crystal produced at room temperature
a (1 1) LEED pattern with sharp spots. The level
of agreement between experimental and calculated I (V )
spectra was confirmed by five metric distances. The best
model showed that the Pt composition oscillates around
the bulk value and damps out over three layers. As
this takes place, the first surface layer is a pure Pt layer
(x1P t = 100 at.%), the second layer is strongly Pt depleted
(x2P t = 48 at.%) and the third layer is only slightly
enriched in Pt (x3P t = 89 at.%). The structural surface
parameters were found to be typical of a (111) face of
an fcc metal, i.e. weak relaxations of interlayer spacings
and
(1d12 = 0.3, 1d23 = 0.9 %, db = 2.226 A)
. . .ABC. . . stacking.
5.1.3. PtNi alloys. Ni and Pt are both transition metals
exhibiting important and specific catalysis properties, and
modification thereof by alloying appeared very attractive.
These two metals after alloying form fcc solid solutions,
which are the subject of high interest in a wide of
fields, particular as catalysts in oxidation, respectively
hydrogenation reactions. On the other hand, these alloys
exhibit some quite remarkable surface features compared
to other alloys with similar electron structure and phase
diagrams, and are probably the most studied among the
surfaces of random substitutional alloys. In the light of our
review this alloy system has aroused considerable interest
because of the tendency to compositional ordering in the
bulk. The equilibrium phase diagram [13] shows that Ni
and Pt atoms form a continuous series of random solid
solution at higher temperature. However there are also
3058

two ordered superstructures: around the composition Ni3 Pt


(Cu3 Au L12 type of structure) with T0b = 853 K, and
NiPt in the equiatomic region (tetragonal CuAu L10 type of
structure) below 918 K; the homogeneity range extends
from about 40 to 55 at.% Pt at 673 K. There is also a
variation of the Curie temperature with the composition of
the alloy [3].
Ptx Ni1x alloy surfaces have been extensively studied
in the disordered state only. Let us observe in brief
terms the most thoroughly studied single-crystal alloys
with following composition: Pt10 Ni90 , Pt25 Ni75 , Pt50 Ni50 ,
Pt78 Ni22 , having different face orientations.
P t x Ni 1x (100). Gauthier et al [127] reported the first
quantitative LEED study to determine both the geometric
structure and the composition of the surface region of the
Pt10 Ni90 (100) alloy. The final surface cleaning procedure
included annealing at 1050 K and then the alloy was cooled
down to room temperature. The electron scattering in the
substitutionally disordered alloy was described by a layerby-layer version of the averaged T -matrix approximation
[127131]. LEED-calculated I (V ) spectra (by the layerdoubling method) and the experimentally observed ones
were compared in those works by means of metric
distances. In the case of the Pt10 Ni90 (100) alloy from such
dynamical LEED analysis a damped oscillatory surface
segregation has been found with a Pt concentration x1P t =
24.3 2.7 at.% in the first layer and x2P t = 6.4 5.9 at.%
in the second surface layer. The deeper layers have been
assumed to be bulklike, i.e. xbP t = 10 at.% Pt. Examine
of the (1 1) LEED pattern showed that a Pt-enriched
outermost layer is not reconstructed. The absence of a
clear (2 2) superstructure which is expected in the case
of compositional order at x1P t = 24.3 at.% Pt and the
presence, on the other hand, of narrow stripes indicating
that [011] rows were partially ordered allowed us to draw
conclusions about the occurrence of small ordered areas. It
was proposed that the tendency of the alloy to order and
thus prefer NiPt bonds against NiNi and PtPt bonds may
be the main reason for the observed oscillation segregation
profile, because it leads to an increase of the number of
NiPt bonds between the outermost layers.
The dynamical LEED analysis also showed an
oscillatory behaviour for the first three interlayer spacings:
the distance between the first and the second layers d12
was expanded by +2.0 0.3% as compared to the bulk
value db ; the distance between the second and the third
layers d23 was contracted by 1.2 0.4%; and the d34
distance was expanded by +1.6 0.9%. It is interesting
that in contrast to Ni(100) and some other fcc (100) metal
surfaces [6] however, the Pt10 Ni90 (100) surface starts with
an expansion of the first interlayer spacing, probably due to
surface segregation of the larger Pt atoms to the outermost
layer.
A strong Pt enrichment in the first layer (up to 80 at.%)
followed by a depletion in the second was also found
in the case of the Pt50 Ni50 (100) alloy by Gauthier and
Baudoing [132] using also quantitative LEED analysis.
It was concluded that this Pt segregation led to the
formation of a (119) surface-reconstructed superstructure.
Recent results by STM relative to the Pt10 Ni90 (100) and

Surface effects

Pt25 Ni75 (100) faces [12] showed that these surfaces, too,
tend to reconstruct forming a shifted row structure with a
locally hexagonal structure.
Ptx Ni1x (110). Gauthier et al [129] using quantitative
LEED analysis have studied the surface structure and
composition of the Pt10 Ni90 (110) alloy.
After Ar
bombardment and annealing at 1200 K the bulk disordered
crystal displayed a LEED pattern with sharp spots typical
of a clean fcc (1 1)-(110) surface. However, within
large energy ranges, halfway between the [100] rows
of the (1 1) structure, dim continuous streaks in the
[100] direction were observed, indicative of a tendency
to (1 2) atomic arrangement. The presence of the
additional weak LEED pattern features were attributed to
two-dimensional composition ordering of the alloy surface.
This partial ordering appeared the most stable situation
for this alloy and was associated with surface segregation
behaviour. Quantitative LEED intensity analysis allowed
us to conclude that the surface concentration and interlayer
spacing oscillate and are attenuated, from the outermost
layer and inwards, over three layers, with final parameters
x1P t = 6.4 3.6, x2P t = 52.3 2.1, x3P t = 10 10 at.% and
(1d12 = 4.5 %), d23 = 1.308
d12 = 1.206 0.09 A

Hence,
0.14 A (1d23 = +3.6 %), d34 = 1.265 0.09 A.
the Pt10 Ni90 (110) face exhibits a well defined NiPt surface
sandwich, where the (110) face favours Ni enrichment.
Gauthier et al [128] also concluded that in addition to a
1D composition ordering perpendicular to the surface there
is a 2D ordering observed parallel to the surface. This
surface ordering tendency for the bulk disordered crystal
was attributed to a partial atomic arrangement in the second
layer with 50 at.% Pt and 50 at.% Ni. As a result Pt and Ni
atoms alternate along the [110] rows so that most Pt sites
have only Ni first neighbours and vice versa. In this case
only one type of Pt catalytic site is present.
The same quantitative LEED analysis was also used
by Gauthier et al [128] to determine the surface layer
composition and multilayer relaxation on the (110) face of
the disordered Pt50 Ni50 alloy. Upon annealing the crystal at
1300 K the surface exhibited a sharp (11) fcc (110) LEED
pattern. It was found that the segregation created a surface
sandwich, which in three layers from the top layer and
inwards contains enrichment of 100 at.% Ni (x1P t = 0 at.%,
x2P t = 95 4 at.% Pt) and 83 at.% Ni (x3P t = 17 7 at.%
and x4P t = 48 13 at.% Pt). The surface segregation
was accompanied by a substantial multilayer relaxation

relative to the (110) spacing in the bulk (db = 1.325 A):


(1d12 = 19%, contraction),
d12 = 1.07 0.01 A
(1d23 = +11%, expansion) and
d23 = 1.47 0.02 A

d34 = 1.31 0.02 A (1d34 = 1%, contraction).


Ptx Ni1x (111). The single-crystal Ptx Ni1x (111) alloy
surfaces have been studied with several experimental
techniques: LEED [130, 131], LEIS [133, 134], XPS [135]
and STM [136]. All the data except STM results were
obtained on crystals which were substitutionally disordered
in the bulk. The first qualitative LEED results obtained by
Gauthier et al [137] and Bertolini et al [138] on (111)oriented faces of PtNi alloys of various bulk composition
annealed above the temperature of orderdisorder transition
always showed at room temperature a (1 1) LEED

Table 2. Pt concentrations in the first and second


outermost layers for Ptx Ni1x (111) alloys [130, 131, 133].

Layer
1
2
3
Bulk

Pt10 Ni90
xiPt (at.%)

Pt50 Ni50
xiPt (at.%)

Pt78 Ni22
xiPt (at.%)

30 4
53

88 2
95
65 10
50

99 1
30 5
87 10
78

10

pattern indicating a complete randomization of the surface


components.
The quantitative crystallographic LEED
analysis of the (111) face of Pt50 Ni50 , Pt78 Ni22 [130]
and Pt50 Ni50 [131] demonstrated in all cases a bulk
termination structure with a strong enrichment in Pt of
the outermost layer and a damped oscillatory composition
profile including three outermost layers (table 2). In
other words, at its surface each one of the studied alloys
segregates into a well developed PtNi sandwich with Pt
atoms on top. Evidence for Pt enrichment in the outermost
layer was also reported by MEIS [134] and XPS [135].
The MEIS study by van de Riet et al [134] was initiated
first to examine the existence of the ordering or clustering
in the Pt50 Ni50 (111) surface, which can be deduced in
principle from the angular position and from the shape of
the intensity increase in a polar angular distribution. It
was found that the Pt50 Ni50 (111) face annealed at 723 K
and cooled down to 353 K contains 88 2 at.% Pt. A
combination of shadowing and blocking techniques and
model calculations demonstrated that the Ni atoms of the
outermost layer are not clustered in domains and gave
no evidence of the surface ordering effects. However,
Schmid et al [136] have performed the first attempt at direct
observation of the surface compositional order by STM.
The sample, a Pt25 Ni75 (111) single crystal, was prepared
by repeated cycles of Ar ion sputtering and annealing at a
final temperature of 780 K. The LEED pattern did not show
any superstructure spots, which might indicate long-range
compositional order; the STM images were compared with
Monte Carlo simulations with embedded atom potentials.
The simulations verified segregation of Pt to the first surface
layer (x1P t = 47 at.%) and Ni enrichment of the second
layer, and showed surface compositional-ordered (1 2)
domains of the kind observed by STM (figure 31). These
domains were close to the Pt50 Ni50 2D ordered phase
because below 918 K the ordered CuAu type L10 phase
exists in the bulk. Hence one can speak of a kind of
surface-induced compositional ordering which is distinct
from the 2D ordered structure expected from the fcc L12
bulk ordered phase.
5.1.4.
Concluding remarks on PtTM alloys. A
number of studies have focused on single-crystal Ptbased alloys, most notably on PtTM surfaces. This
is because some of them have interesting catalytic or
chemical properties directly connected with the strong Pt
surface segregation [128131, 138]. At the present day
the quantitative full LEED intensity analysis developed by
Grenoble group [132] appears to be the powerful method
3059

M A Vasiliev

Figure 31. Atomic arrangement on the Pt37 Ni63 (111)


surface at 420 K calculated by Monte Carlo simulations
[136].

to give important information on the PtTM (TMNi, Co,


Fe) surface region with layer-by-layer resolution. It is
possible to compare results obtained on examples of Pt
TM alloys having the same face orientation and the same
bulk composition, namely, Pt78 Ni22 (111), Pt80 Co20 (111)
and Pt80 Fe20 (111) alloys [11]. All these three alloys are
fcc with the usual unperturbed ABC stacking sequence of
(111) planes. However there are also some differences: first
the pure transition metals alloyed with Pt have different
bulk lattices, i.e. Ni is fcc, Co is hcp and Fe is bcc;
second, the orderdisorder critical temperature for ordered
Ni0.5 Pt0.5 (915 K) and ordered Pt3 Co (1098 K) phases are
substantially below than that for ordered Pt3 Fe (1623 K)
phase. Because of this it was impossible to achieve clean
surfaces for the ordered state in PtNi and PtCo alloys
in contrast to PtFe alloys. In the case of the (111)
face the Grenoble group has established that the outermost
layer is essentially of quasi-pure Pt for all three abovementioned alloys, and the composition profile damps out
to the bulk value within four outermost layers. However
the deeper layers showed different behaviour depending on
alloy: marked Pt damped concentration oscillation for Pt
Ni and PtCo alloys (but less pronounced) and a monotonic
decrease for the PtFe ordered system (figure 32). In the
last case, additionally the compositional order was observed
in the underlying layers. It was concluded that the change in
the compositional profile (pronounced oscillation for PtNi,
monotonic profile for PtFe and intermediate situation for
PtCo) is the result of balance between the strain energy
term, caused by differences in the atom size and lattice
parameters, and the ordering energy term, i.e. strong Pt
TM pair interaction in the ordering systems which is very
effective to stop the oscillation behaviour.
5.2. Compound-forming alloys
5.2.1. CuAl -phase. AlCu alloys are a well known
constructional material in varied engineering applications
owing to a good combination of strength and toughness
at lower densities. Al and Cu atoms form fcc random
substitutional alloys in a range of composition from 0
to 19.6 at.% Al (-phase). Above this range several
ordered compounds exist [13]. The surface of this alloy
3060

Figure 32. The layer dependence of the Pt surface


concentration for PtM(111) alloy (M = Fe, Co, Ni).

has attracted particular interest because the (111)-oriented


face in the case of -phase was found to be reconstructed
[57, 139142], giving an example of an alloy surface which
exhibits LRO at the surface in the absence of LRO in the
bulk. Ferrante [139, 140] first demonstrated a sharp well
defined LEED pattern for
the Cu10
at.% Al(111) face,
which was interpreted as ( 3 3)R30 superstructure.
Using AES he also indicated increased Al concentration
was limited to two layers up to 43 at.%. Later, this
surface

superstructure was shown to undergo a reversible


( 3 3)R30
(1 1) phase transformation at 570 K
on the (111) face of Cu12.5 at.% Al alloy [57]. This
transformation was also observed at approximately the same
temperature in the Cu16 at.% Al crystal by Baird et al
[142].
At first, it was suggested [57, 140, 141] that the

( 3 3)R30 superlattice was due to an ordering of


Al atoms above the alloy surface. But this Al overlayer
model was not supported by Baird et al [142], using the
quantitative LEED intensity analysis. Since this work has
given evidence of a radically new structure model we
will look at it more closely. The Cu16 at.% Al (111)
face was cleaned by Ar ion bombardment
at 500 eV
and after annealing below 570 K the ( 3 3)R30
superstructure was observed. In this case the surface
concentration of Al, as determined by AES, had increased
to one-third of a monolayer. Once the superstructure had
formed, the AES peaks did not change as the specimen
was heated. At 570 K the reconstructed surface exhibited a
reversible orderdisorder phase transformation to (1 1)
symmetry. I (V ) spectra were recorded for the ( 3

3)R30 superlattice with the crystal cooled to 150 K.


For the LEED intensity calculations well established multiscattering schemes were used [57]. Comparison between
theory and experiment was made by the five type of Rfactor.
The four structural models were calculated for the

( 3 3)R30 superstructure, including (i) one-third of a


monolayer of Al located in both fcc and hcp hollow sites
on pure Cu(111), and (ii) one-third of a monolayer of Al
located substitutionally within the top layer of fcc Cu(111).
In addition the mixed top layer in the second model was
this means that
given bucklings of 0.0 0.1, and 0.2 A;
Al atoms were depressed below or raised above the top Cu
layer by the amounts indicated. Also, the interlayer spacing

Surface effects

Figure 33. Top view of the -CuAl(111) surface [142].

between the mixed top layer and the next pure Cu layer was
The best agreement between
varied from 1.887 to 2.287 A.
theoretical and experimental I (V ) spectra was consistent
with the above-mentioned second model, but with a slight
tendency toward a small expansion of the top layer spacing
and a small inward buckling of the Al atoms
by +0.05 A
(figure 33). Baird et al [142] have drawn also
by 0.025 A
the conclusion that such a substitutionally ordered surface
was consistent with the short-range order observed in bulk
-AlCu compound-forming alloys.
5.2.2. Al6.5 at.% Li. Considerably recent attention
has been focused in particular on AlLi alloys due to
advantages of these materials in different engineering
applications, for example, as contemplated aerospace
materials, as an anode material in high-performance
secondary cells and lastly as a potential first-wall material
for Tokamak reactors.
However AlLi alloys suffer
from their relatively low resistance to corrosion by the
atmosphere and by water. The later has given rise to
study surface properties of this alloys. According to the
phase diagram in this system a eutectic exists between an
Al-rich random solid solution and a compound, Li2 Al3 or
more likely LiAl, with a bcc structure of the CsCl (B2)
type [13]. It is not surprising, then, that the prediction
of the surface structure model and surface composition
of the AlLi alloys is rather difficult if not impossible
to realize. The first experimental study devoted to this
problem has been performed recently by Esposto et al [143]
using LEED, AES, SIMS and work function measurements.
They investigated Al6.5 at.% Li sheet with facets on the
surface of (111) orientation. The surface was cleaned by Ar
ion sputtering at temperatures between 300 and 600 K. The
LEED pattern observed after the cleaning procedure always
indicated that the surface consisted of (111) facets only.
LEED measurements were carried out during observed Li
segregation (figure 34). It was noted that rapid segregation
of Li atoms to the surface layers (measured by AES) took
place at annealing above 500 K, reaching a maximum Li
composition of 0.17. At higher temperature the Li surface
concentration decreased because of evaporation. It was
found that at annealing temperature 500K the
LEED
data indicated the formation of an ordered ( 3 3)R30
superstructure on the substrate with hexagonal symmetry.
Upon cooling the sample, the value of Li segregation
remained on the surface resulting in a final average surface
concentration of 0.16. The work function showed that a
rapid enhancement of surface Li concentration associated
with a fast decrease in 1 in the same temperature range.

Figure 34. The temperature dependence of the Li surface


concentration for Al6.5 at.% Li [143].

6. Comparison of theoretical and experimental


studies
The present review is mainly concerned with experimental
studies of the various surface effects in ordering alloys. On
the other hand, many interesting theoretical results with
surface orderdisorder phenomena were obtained over a
period of years. A review of these is not the aim of
the present paper. Nevertheless we will focus attention on
some related theories in the discussion of experiments and
confine ourselves for the present to recalling some basic
aspects giving surface characteristics of the real ordered
alloy systems.
6.1. The surface orderdisorder phenomena
It was first shown theoretically by Valenta and Sukiennicki
[8] for a thin film of Ni3 Fe with (111) orientation as an
example, that changes in the LRO are to be expected at
the surface of ordered alloys. They and then Sukiennicki
[144] have surmised these changes must be a result of a
natural surface defect, i.e. lack of a fraction of nearest
neighbours for the surface atoms, and also different lattice
constant and interlayer spacing in the surface region. In
general, all kinds of surface defect, connected with different
conditions of surface layers in comparison with those in
the bulk make their contribution to the surface order
disorder behaviour. Sukiennicki and Puszkarski [145] have
proposed a phenomenological surface parameter as a
measure of the change in the ratio of the ordering energies
due to surface defects for surface atoms and that for inside
the bulk. The ordering energy as it usually in the case of the
interaction between nearest neighbours only was defined
V = EAB 12 (EAA + EBB )

(5)

where EAB , EAA and EBB are the pairwise energies


of interaction between the AB, AA and BB atoms
respectively for Ax B1x alloy. From the condition for
thermodynamic equilibrium at a given temperature,
F
=0
i

i = 1, 2, . . . , n

(6)

where F is the free energy of a system, and i is an LRO


parameter for the ith monolayer. The authors of [145] used
3061

M A Vasiliev

6.2. Ordering and surface segregation

Figure 35. Comparison between calculated ([10]) and


measured ([10], [17]) LRO parameters and Au
surface concentration, calculated ([32]) and measured
([15]), for Cu3 Au.

the well known BraggWilliams approximation and made


numerical calculations of i for several values of . In
doing so they considered the first two boundary surface
layers only. The results of calculation for Ni3 Fe(111) are
shown in figure 23 in comparison with experimental data
by Vasiliev et al [146] considering the surface segregation
effects. In view of this theoretical work one can consider
that change of the surface composition is also a surface
defect which influences both and i parameters. As we
can see from figure 23 the surface values of i according
to the experimental and theoretical results are lower in a
comparison with b at all temperatures. The theoretical
data also show different behaviour of the surface parameter
i depending on parameter ( = 1 is the case of the
natural defect only). In particular if is large than 3/2
(critical value) the parameter i may be even higher than
the bulk one, that is not the case for Ni3 Fe.
The behaviour of the LRO parameter as a function
of temperature for ordered Cu3 Al(100) alloy was first
calculated by Sundaram et al [10] also on the assumption
that at each temperature the composition of the layers
corresponded to a bulk layer. Nonetheless, they have
obtained some fundamentally new theoretical as well as
experimental results for ordered system. In the statistical
computations of the i parameter in layers near the surface
the Monte Carlo method was used, and interactions between
the atoms in Cu3 Au were represented by an Ising-like
model Hamiltonian. The result of calculations for the real
mixed CuAu termination of the surface model is shown
in figure 35 together with LEED data. It can see that
theoretical data are in reasonably good agreement with the
experimentally deduced order parameter which, unlike the
bulk, appears to be a continuous function of temperature
for the surface. This was new evidence of the difference
between observed behaviour of the LRO parameter at the
alloy surface and that in the bulk. However the influence
of the surface composition as a function of the temperature,
which was later experimentally observed, was not predicted
in the calculations.
3062

The theory of surface segregation in an ordered alloy


was first developed by van Santen and Sachtler [85] for
the Pt3 Sn compound. They used the BraggWilliams
approximation in order to evaluate the surface free energy
in both the ordered and disordered state assuming the same
order parameter at the surface and the bulk crystal. This, of
course, is not correct as was shown above. Nevertheless, it
was demonstrated that the preferential termination of the
(100) Pt3 Sn crystal in the compositionally mixed (100)
plane is the result of the surface composition constrained
by strong ordering energies in qualitative agreement with
experimental observation [86].
In succeeding years the A3 B ordered system (fcc,
Cu3 Au type) has been the subject of several theoretical
investigations of the surface segregation effects on the
orderdisorder behaviour at the surface. This problem was
raised by Moran-Lopez and Bennemann in their pioneering
work [148]. The orderdisorder phase transition and
surface composition for A3 B type alloys have been studied
using the BraggWilliams approximation.
If surface
segregation was neglected, the calculated temperature
dependence of the surface LRO parameter accorded well
with theoretical result reported earlier for Cu3 Au [10].
When the effect of surface composition changing on
ordering was taken into account the applied theory predicted
that the order parameter in the vicinity of the surface
can be strongly reduced and might even disappear at the
surface at some temperature below the bulk transition
temperature T0b (see the above example of Ni3 Fe). It
was the first theoretical prediction indicating that surface
segregation and composition ordering in alloys affect each
other. Experimental support for this important conclusion
was obtained more recently when the experimental LEIS
data were reported [15].
Using the cluster-variation method in the tetrahedron
approximation Kumar and Bennemann [147] have studied
the orderdisorder transition at the (100) surface of the
Cu3 Au alloy considering the experimental results of the
temperature-dependent surface segregation [15]. Their
calculations are evidence in favour of a general conclusion
made previously [148] that there is a competition between
segregation and ordering. For example, depending on
surface segregation behaviours as function of temperature
the surface orderdisorder transition may be first or second
order (the case of Cu3 Au), and disordered phase at the
surface may occur at temperature above or below the bulk
transition temperature (the case of Ni3 Fe). The calculated
surface LRO agreed with the LEED data [10] but the Au
surface concentration in Cu3 Au(100) was far below the
LEIS result for both ordered and disordered structures.
The above-mentioned theoretical studies had the grave
practical disadvantage that the simple model involved only
atom interactions which were the same for both surface
region and bulk crystal. In the subsequent study Sanchez
and Moran-Lopez [32, 54] demonstrated that the surface
orderdisorder effects obtained for the (100) surface of
Cu3 Au can be in fact accurately treated by a simple
Ising model with nearest-neighbour pair interactions in
which such interactions at the surface were assumed to

Surface effects

Figure 36. The layer dependence of the Au concentration


calculated for Cu3 Au(100) at 670 K [32].

be different from those in the bulk. The free energy


of the system at finite temperatures was calculated by
use of the self-consistent cluster-variation method in the
tetrahedron approximation. As seen in figure 35 the
agreement between the model presented by the authors of
[32, 54] and the available experimental LEED and LEIS
results is satisfactory both in considering to the surface
Au segregation (x1Au , x2Au ) and to the continuous surface
disordering occurring at the same critical temperatures
as the bulk first-order transition, i.e. T0s = T0b . The
Au concentration and the LRO parameter have been also
calculated for the 20 planes adjacent to the surface and so
the Au layer-by-layer profile was obtained at a temperature
of 670 K, 7 K above the T0b (see figure 36). One can
see from this figure that such an oscillatory concentration
profile in a direction perpendicular to the surface is an
inherent feature of ordering alloys. As we can see in
figure 35 the theoretical surface layer concentration shows
a discontinuous jump at the T0b in contrast to the LRO
behaviour. The authors of [32] concluded that this effect
is dictated by the first-order discontinuities in the bulk
LRO parameter. We leave here the discussion of the
interplay between surface segregation and ordering for the
Cu3 Au alloy and propose to the reader who has a taken
an interest in this question the papers [35, 149155] which
also gave the theoretical treatment of the Cu3 Au ordered
alloy surface.
Let us consider the NiPt alloy system next most
accepted for theoretical treatment, but only from the
viewpoint of ordering effects compared with experimental
data. The strong Pt (or Ni) surface segregation and
its extreme sensitivity to the face orientation in NiPt
ordering alloy have elicited a number of theoretical models
[29, 66, 128, 153, 156163] in order to predict the relative
stability of different possible configurations, such as the
observed sandwich and the segregation inversion in (110)
surface. It must be emphasized that the NiPt alloy
surfaces, notably the (110) surface, are found to present
some problem for the calculations. The first problem is
connected with some factor which makes it preferable for
Pt atoms to be situated at (100), (111) surfaces and Ni
atoms at (110) surfaces; the second problem has motivated
the nature of the strong depthcompositional oscillation.
The first theoretical estimations have shown that the
commonly used simple bond-breaking models [75, 164] fail

to predict Pt segregation in the case of NiPt alloys. It


is well known when the two kinds of atom in an alloy
rP t = 1.39 A,
for
have different radii (rN i = 1.24 A,
example), there is a strain in the lattice (size effects). This
effect can be calculated by means of elastic continuum
theory [165]. However Lundberg [159] has demonstrated
that surface Pt segregation was poorly described by the
bond-breaking model with the elastic strain term [165].
The conclusion is that these simple theories including the
heats of vaporization of pure metals, surface tension (quite
negligible), size effect and the bulk activity coefficients
are inapplicable to NiPt alloy surfaces. The best results
were obtained indicating a non-regular behaviour of the
NiPt system, which as was pointed above has a strong
ordering tendency towards the formation of ordered L10
and L12 phases. For this reason the most theoretical
studies have taken into account some ordering effects
[29, 153, 156, 157, 160163]. Unfortunately, it is almost
impossible to test these effects for well ordered phases
because all experimental studies of Pt and Ni surface
segregation for all concentrations and faces were performed
with random alloys. In spite of this fact the incorporation
of the ordering term in some theoretical models has allowed
us to obtain a number of fundamental data. From this
point of view let us briefly consider some of the works
which used the most developed theoretical model, the tightbinding Ising model (TBIM) [29, 159, 166].
Treglia and Legrand [166], Lundberg [159] and
Legrand and Treglia [29] have shown that the main
features of the surface segregation phenomena for Ni
Pt alloys can be interpreted using an improved TBIM
approximation in which the size effect was treated within a
microscopic model and the pair interaction V s was varied
at the surface region according to the electron structure
calculations. In particular it has been shown that the
sign of V (V > 0 in equation (5) is a tendency to
ordering) is due to the spinorbit coupling interactions. The
V -value was also derived from the experimental critical
temperatures of the L10 ordered phase in the mean-field
approximation (V b = 12 kT0b = 0.038 eV). However for
different crystallographic orientations the enhancements at
the surface of the effective pair interactions are increased
when the face is an open one, mainly V b < V s (111) <
V s (100) < V s (110). For example, in the last case
1.5 < V s (110)/V b < 2. The different V s values were
involved in the segregation energy calculations in order to
determine the physical origin of this phenomenon. Finally
the segregation energy was reduced to the combination of
the two main factors: of the size mismatch term (1Hise )
and the ordering term 1H ord . It was demonstrated by
TBIM calculations that the competition between 1Hise
and 1H ord terms may be thought of as the genesis of a
variety of segregation behaviours in relation to the surface
orientation, the bulk concentration and the temperature
region.
This competition led to a strongly damped
oscillating segregation profile at T > T0b for the (100)
and (110) surfaces compared to the (111) surface, which
was explained as a tendency towards the formation of
the ordered phases in the bulk. The size effect appeared
more pronounced on the closer-packed (100) face than on
3063

M A Vasiliev

the more open (110) surface. To summarize, the TBIM


model for the (100) and (111) surfaces predicted a Pt
surface enrichment for the whole bulk concentration range
in good agreement with experimental results obtained by
LEED. Nevertheless the discrepancy remains still in a large
difference between the LEED and TBIM for the (110)
surfaces. Finally, experiments in the ordered bulk phase of
the NiPt alloys should be of great interest because they
may give a clear insight into the nature of the relation
between surface segregation and ordering effects in both
ordered and disordered NiPt alloys.
In conclusion it may be pointed that the magnetic and
vibrational contributions are small in NiPt alloys and the
entropy of alloy formation 1S f orm in the Pt-rich region
suggests a tendency towards ordering [167]. On the other
hand, Pt50 Ni50 (110) alloy is non-magnetic in the bulk but
the question now arises of whether sufficient spin coupling
would occur among the Ni atoms at the (110) surface for the
observed NiPtNi sandwich to exhibit surface magnetism
[128].
Now we note two examples when the improved
quasichemical broken-bond model was usefully employed
for the theoretical prediction of the surface segregation in
the strongly ordered intermetallic compound Pt3 Ti [168]
and ordering Cu85 Pd15 alloy [94]. This qualitative model
of segregation [164] gives two simple criteria for predicting
the general segregation behaviour in very dilute ideal alloys:
(i) the bond strength ratio, (ii) the atomic size ratio. Spencer
[168] has shown that when these criteria are applied to
dilute solid solutions of Ti in Pt the structure segregation of
Ti should occur in contrast to the experimental results [84].
However he extended this model to a more concentrated
alloy, in particular the ordered Pt3 Ti compound, and used
thermodynamic data for PtTi alloys as well as those for the
pure metals. Considering the PtTi alloys as the irregular
system (ordering) Spencer [168] proposed, on the basis of
the broken-bond model, that two main factors control the
thermodynamics of segregation: the strain energy and the
surface free energy. The first factor as known originates
from the mismatch of the sizes of the atoms, and the second
one comes from bonding between atoms. In the pure metals
rP t =
Ti has the larger metallic radius (r T i = 1.47 A,

1.39 A); however for the Pt3 Ti the effective size of the Ti
atom appeared to be smaller than that of the Pt atom [164]
so that factor does not favour the segregation of Ti atoms.
Moreover for Pt3 Ti any TiTi bonding can be neglected
because all Ti atoms are coordinated to 12 Pt atoms as
nearest neighbours while the Pt atoms are coordinated to
eight Pt atoms and four Ti atoms as nearest neighbours. In
this case the free energy of segregation
1Gseg = 1Z(EP tT i EP tP t )

(7)

consists only of the nearest-neighbour bond strengths


EP tT i and EP tP t ; 1Z is the difference in effective
coordination number of a bulk site and a surface site.
Since the free energy of the PtNi bond (EP tN i =
22.17 kcal/g-atom) is larger than the PtPt bond (EP tP t =
17.7 kcal/g-atom) at 800 K, the surface free energy will
also not allow Ti atoms to occupy the surface, in line with
experiment [84].
3064

Newton et al [94] have also used the simple brokenbond model to predict surface segregation in the CuPd
ordering alloys despite the fact that this model is not
formally designed for systems having a tendency to order
[75, 164, 169]. The results of these calculations have shown
in agreement with experimental observation a Cu outermost
layer, and a second layer significantly enriched in Pd for
the Cu85 Pd15 (110) equilibrium structure. It was again
demonstrated that the overall driving force for segregation
in ordering alloys (having a negative heat of solution) is
the result of a competition between the size mismatch and
surface free energy terms.
Thecondition for orderdisorder transition to the
( 3 3)R30 structure ordered phase in the absence
of the bulk LRO was discussed by Teraoka [170] using
of the BraggWilliams approximation. In fact such a case
was experimentally observed on an fcc (111) surface of
-AlCu alloy with bulk Al concentration of more than
9 at.% [140, 142]. Using the nearest-neighbour pairwise
interaction energies Teraoka defined the ordering energy
V = EAA + EBB 2AB and the surface segregation energy
W = E
AA EBB for the Ax B1x (111) surface and ordered
( 3 3)R30 phase. The role of the surface segregation
in the surface orderdisorder
transition in order to form the
outermost layer with ( 3 3)R30 symmetry was shown
first. This structure was predicted for the absence of the
bulk LRO only if W/V > 1 and x < 0.21 by calculations
of the difference between the free energies of the ordered
and disordered phases [170].
6.3. Surface magnetism
It is well known that the compositional order of atoms
may significantly influence the bulk magnetic properties
of alloys [3]. So in a number of ferromagnetic alloys the
saturation magnetization, Curie temperature, coercive force,
magnetic permeability and some other properties are found
to be strongly dependent on the degree of compositional
order.
There are a number of theoretical analyses of the
interdependence of magnetism and order in the bulk
binary alloys with one or two ferromagnetic components
[4, 5, 171173]. In particular the influence of magnetism
on orderdisorder critical temperature and, conversely,
the influence of ordering on the Curie temperature were
treated in detail for the bcc Cox Fe1x alloy [4, 5]. It
was also shown that the interplay of compositional
and magnetic order leads to complicated bulk phase
diagrams which can only be obtained if both effects
are considered on an equal footing [4, 5, 174]. Based
on a study of the bulk properties one may speculate
that the situation at the surface of the ferromagnetic and
compositionally ordered alloys must be more complicated.
Firstly this is due to the well established fact that at
least for ferromagnetic pure metals (Fe, Co, Ni, Gd)
the magnetization at the surface may differ considerably
from that in bulk crystal because of the reduction in
the coordination of the surface atoms as well as the
dependence of magnetic surface properties on the chemical
state of the pure metal surfaces (magnetochemistry effect).

Surface effects

More complete reviews on surface magnetism of pure


metals have been given by Mathon [175] and recently
Gradmann [176, 177].
It is worth noting the main
fundamental advantages in surface magnetism study in
terms of our review are the following. (i) Powerful
experimental techniques for surface magnetic analysis in
UHV were developed, in particular SPLEED, ultrathinfilm spectroscopy, Mossbauer spectroscopy and electroncapture spectroscopy. (ii) On theoretical grounds, surface
magnetism has been treated within different frameworks,
in particular the mean-field approximation, effective
field theories, the random-phase approximation, Monte
Carlo techniques and renormalization group methods.
(iii) The first-principles calculations of band structures,
magnetic moment and hyperfine field near surfaces as
well as experimental studies have confirmed that magnetic
moments are enhanced in free surfaces of transition
metal ferromagnets and oscillation of hyperfine fields; the
enhanced surface magnetization penetrates some distance
into the bulk. (iv) The temperature dependence of surface
magnetization is basically understood: the magnetization
at the surface decreases more rapidly than the bulk
magnetization as the temperature is raised. (v) Magnetic
surface anisotropy as a result of broken local symmetry
in surfaces has also now generally been accepted as an
important basic surface property. (vi) Modified exchange
interaction Iis near the free surface gives rise to some
type of surface magnetic order even above the bulk Curie
temperature, so-called live (surface) layers.
It is necessary to stress that the nature of the phase
transitions in magnetic systems has been studied extensively
both theoretically for an infinite system [47, 178184] and
experimentally for finite systems with real surfaces (see
reviews [175177]). In terms of our paper it is worth
noting that the surface and bulk magnetization curves as
functions of the temperature depend on how Iis /I b varies
in accordance with theoretical prediction. As a result the
phase diagram for the semi-infinite Ising model depending
on the values of Iis and I b was proposed [183]. As indicated
in figure 37 depending on the value of Iis /I b different
behaviours at the surface and bulk can be obtained. If Iis is
s
smaller than a critical value Ii,C
(T < TCb ) both the bulk and
the surface are magnetically ordered (ordinary transition);
s
and if Iis > Ii,C
, the surface is disordered (paramagnetic
phase) at a temperature Iis (surface transition) larger than
the bulk TCb (extraordinary transition), i.e. the surface
remains magnetically ordered while the bulk order is absent.
In the last case for the intermediate values of TCb < T < TCs ,
the magnetization decays exponentially into the bulk with
characteristic length [183].
6.4. Surface effects in magnetic ordered alloys
Of special current interest is the magnetism of alloy
surfaces, in particular the relation between the atomic
structure, surface segregation, compositional and magnetic
order in the surface region. So far this problem is not
well understood in terms of either theory or experiment as
compared with surface magnetism of the single-component
metals. In the case of alloys the interplay of physical

Figure 37. Phase diagram for the semi-infinite Ising model


with bulk coupling J b and surface coupling J s . The state of
the bulk is denoted BF for a ferromagnet and BP for a
paramagnet; the surface phases are SF (ferromagnet), SP
(paramagnet) and SAF (antiferromagnet) [183].

processes is more complicated because more surface and


bulk parameters must be included in the theoretical model.
On the other hand the experimental study of alloy surface
magnetism is also very difficult because there are not so
far experimental techniques which allow reliable data to
be obtained over the wide range of temperatures under
the UHV test conditions making possible in situ surface
analysis of the atomic structure, composition and magnetic
properties in a wide surface region. Nevertheless, a
few theoretical and experimental results were reported
which may be of interest in future study of alloy surface
magnetism. From this viewpoint more interesting systems
are those alloys in which compositional and magnetic order
occur simultaneously.
In the theoretical studies by Moran-Lopez and Falicov
[185, 186] and Urias and Moran-Lopez [187] it has been
first shown that in comparison with the pure metals
the surface magnetization in alloys differs significantly
from that in the bulk not only owing to the diminishing
of the surface coordination number but also because of
differences in the surface composition and degree of
compositional order. For example, the surface segregation
of ferromagnetic components in the dilute alloy with one
magnetic component may lead to a surface magnetically
harder than the bulk according to theoretical prediction
[188, 189] and experimental observation, particularly in the
case of random NiPt alloys (see above). The effects
of ordering we illustrate by the example of the ordered
ferromagnetic bcc Fe0.5 Co0.5 alloy which was understood
by both experimental and theoretical studies [47, 51, 185
187, 190].
The first theory of the relation between surface
composition, short-range order parameter and magnetism
for binary bcc Ax B1x applied to ordered Fe0.5 Co0.5 (110)
was initiated by Moran-Lopez and Falicov [185, 186]. To
calculate the surface properties of such a ferromagnetic
system they used a Heisenberg Hamiltonian with
3065

M A Vasiliev

interactions between nearest neighbours only, and a


surface structural model consisting of a mixed Bethe
lattice. In this model the surface topmost layer was
different from the bulk planes only. The temperature
dependence of the surface composition, the short-order
parameter, the spin-wave spectra, the bulk and local surface
magnetization were calculated with parameter fitted to the
well known experimental data (Curie temperature, heat of
vaporization and orderdisorder temperature). The surface
Fe concentration as a main result appeared to be higher than
in the bulk and the spins predicted were SF e = 2 > SCo =
1.5. As consequence in the case of the Fe0.5 Co0.5 (110)
ordered alloy the surface exhibited a large magnetization
compared to the bulk for temperatures observed up to
400 K. On the other words the atoms with larger spin were
segregated to the surface.
In succeeding years Urias and Moran-Lopez [187] for
the same system have considered the compositional LRO
with surface segregation and magnetic order parameters.
The calculations were performed within the mean-field
(BraggWilliams) approximation pairwise between nearest
neighbours only. The magnetic interactions were surmised
to be Ising type also with nearest neighbours. Let us
underline here only some of the main features of this study,
particularly for the Fe0.5 Co0.5 (110) ordered alloy of the bcc
Ax B1x system with spin SA and SB .
In calculations [187] the heat of alloy mixing was
defined by
W = WC + WM
(8)
where WC = EAA + EBB 2EAB , and WM = 2SA SB JAB
SA2 JAA SB2 JBB are the chemical (ordering term) and
magnetic contributions to the heat of alloy mixing,
respectively. The main driving parameter was defined as
1=

1C WC 1M WM
WC + WM

(9)

EAA EBB
WC

(10)

SA2 JAA SB2 JBB


.
WM

(11)

where
1C =
and
1M =

magnetic terms for the surface and bulk region, as well


as the temperature dependence x1F e (T ) are illustrated in
figure 38. One can see in the figure that the surface
LRO parameter 1s , the magnetic order parameter 1s and
the surface Fe composition x1F e are different from these
in the bulk of primarily completely ordered Fe0.5 Co0.5
alloy. Unfortunately the experimental data are only for
surface composition from AES analysis [47, 51] which is in
qualitative agreement with the discussed theoretical result.
Because the experimental evidence of Gradmann
et al [191] and Victora [192] has confirmed that the
magnetization of the pure free Fe(110) surface is enhanced
by approximately 30% one can conclude that the Fe
segregation in the FeCo system also enhances the surface
magnetic moment. On this basis we may suggest that in
the FeNi alloys (see above) the surface magnetization is
also enhanced due to the strong surface segregation of Fe
observed experimentally. However this alloy system was
not theoretically treated in view of the surface magnetism
phenomena. The depth distribution of the magnetic moment
in the surface region of both ordered and disordered alloys
is also an open problem.
7. Conclusions

The equilibrium values of calculated parameters were


determined by minimizing the free energy. It was first
found that for a bcc Ax B1x system with spins SA ,
SB and for different cases of the relation between Ising
exchange integrals JAA , JAB , JBA and JBB the magnetism
at LRO parameter = 0 favours surface enrichment of the
component A if SA2 JAA < SB2 JBB .
The Fe0.5 Co0.5 alloy is a case with T0b (1003 K) < TCb
(1258 K) and a positive magnetic ordering energy (1M >
0), but with chemical term 1C > 1M . The theoretical
calculations were performed for parameters JF eF e , JCoCo ,
JF eCo determined from the TCb at xFb e = 0.5, and EF eF e ,
ECoCo , EF eCo determined from the cohesive energy of the
pure elements. It turns out that for Fe0.5 Co0.5 (110) alloy the
chemical effects would tend to segregate the Fe atoms in
contrast to magnetism predicting Co surface segregation.
The interplay of the contribution of both chemical and
3066

Figure 38. The temperature dependence of the Fe surface


concentration, bulk and surface LRO parameters and
average magnetization calculated for the Fe0.5 Co0.5 system
[187]. The experimental result [47] is shown by the bar.

Let us summarize briefly the most important achievements


in the studies of the ordering and magnetic effects at the
surface region of the single-crystal binary alloys.
First of all a number of analytical methods very
sensitive to the surface region has been developed to
obtain a more detailed knowledge of the crystallographic
symmetry, lattice and compositional order parameters as
well as the composition and spin order of the outermost few
atom layers of alloys with ordered or disordered (random)
phases. Among the well known methods (listed in the
introduction) some of them were significantly improved
specially for study of surface ordering and magnetic effects
in alloys. In particular the quantitative LEED intensity
analysis is one of very few techniques capable of resolving,
layerwise, the surface structure, order parameter and
composition profile of substitutionally ordered or disordered
phases. In addition, quantitative observation of the surface
orderdisorder kinetics as function of a temperature and

Surface effects

depth was also made possible by the LEED system with


a position-sensitive detector. The combination of LEED
and LEIS (with the time-of-flight version) has a significant
place in the structure and composition analysis of the first
two atom layers. The first STM study was presented which
has demonstrated the clear discrimination of two chemical
species in an alloy surface and short-range ordering effects.
The important result obtained by SPLEED in the study
of the surface magnetization in the ferromagnetic surface
was also shown. SPLEED will remain in the future the
most accurate method for absolute measurements of the
surface magnetization. Using LEED with modulated beam
intensity allowed us to observe some new ferromagnetic
anomalies in alloys. However the lack of extensive and
accurate studies of the alloys surface magnetic properties
in comparison with pure metals is certainly dictated by the
specific experimental difficulties which may be resolved in
the immediate future. Along this line the use of dynamical
SPLEED theory and the layer-by-layer approach are more
appropriate to obtain insight into the atomic magnetic
moment distribution as a function of the temperature and
depth in combination with depth profiling of the lattice and
order parameters as well as layer composition.
Of main interest is the following phenomenon observed
from the studies of the various ordering alloys.
An ideal bulklike termination with (1 1) structure
and compositional mixed (100) layer for the h100i
orientation of the fcc-ordered Cu3 Au, Cu3 Pt, Ni3 Fe alloys
and ordered compounds Ni3 Al, Pt3 Ti, Pt3 Sn as well
as bcc-ordered Co0.5 Fe0.5 (100) alloy and Al0.5 Ni0.5 (110),
Fe3 Al(110) compounds.
Surface segregation damped oscillatory profile in
ordering alloys as systematically found for Ptx Ni1x
random systems and in the disordered Cu3 Au and Ni3 Fe.
Face-related sandwich segregation, observed on
Ptx Ni1x (110) versus (111).
Strong surface multilayer and rippled relaxation as
well as buckling in the first few layers of an ordered
compound (Ni3 Al).
The surface reconstruction with layer relaxation
and buckling, based on the formation of the surface
superstructures of different types, in particular missing
row
(1

2)
reconstruction
in
Pt
Fe
(110);
hexagonal
(
3
80
20

3)R30 in ordered Pt3 Sn(111) and random compoundforming AlLi and AlCu alloys; a quasipure Pt top layer
in ordered Pt3 Ti(111).
The surface order parameters, in contrast to the
bulk behaviour, appeared to be a continuous function of
temperature (second-order transition) in the Cu3 Au and
Ni3 Fe alloys with the bulk first-order transition. It is also
an important observation that the ordering process at the
surface of the Cu3 Au, Ni3 Fe begins at a temperature below
that in the bulk, however in the first case T0s = T0b , and
in the second case T0s < T0b . From layer to layer the kind
of orderdisorder transition in these systems changes from
second order to first order.
In the Ni3 Fe(100), (110), (111) alloys surface
disordering is accompanied by a strong temperaturedependent Fe surface segregation.

The effects of ordering on the surface electronic


structure were observed in detail on an example of ordered
and disordered Cu3 Au(100) and (110) surfaces.
Systematic studies of the surface behaviour of
ferromagnetic single-crystal alloys have only recently
been performed on the example of ordered (disordered)
Ni3 Fe(111). Since its surface magnetic and electronic
structure and composition appeared to be different from
those of the bulk, a number of new surface magnetic
effects have been discovered, in particular, the shift near
the Curie temperature of the LEED reflex energy maximum
and nonmonotonic temperature-dependent behaviour of this
maximum. In addition the antiferromagnetic coupling
between the surface layer and underlying bulk in
the Ni3 Fe(111) alloy was revealed; the surface Curie
temperature appeared to be higher than the bulk one,
because of strong Fe surface enrichment with temperature
increasing.
A number of theoretical models have been developed
specifically to describe the surface behaviour in nonmagnetic and ferromagnetic ordered (ordering) alloys.
The modern realistic models are based on two important
concepts: (i) the pairwise interaction energies in the
surface region are different from those in the other
region, and (ii) the multilayer (layer-by-layer) approach
is far more appropriate to predict the distribution of the
different structural, compositional as well as other physical
parameters as a function of the depth, including order
parameters and kinetics of the surface phase transition.
The interrelation between the compositional ordering,
magnetization and surface segregation was confirmed with
a good agreement with experimental data.
Finally it may be said that the ordering and magnetism
at the surface of the alloys with transition metals will clearly
remain an area of intense activity.
Acknowledgments
The research described in this publication was made
possible in part by grant NU8200 of the Ukrainian
Government and International Science Foundation. The
author is very grateful to Klaus Wandelt for stimulating
discussion and for reviewing the paper.
References
[1] Krivoglaz M A and Smirnov A 1964 The Theory of
OrderDisorder in Alloys (London: McDonald)
[2] Ducastelle F 1991 Order and Phase Stability in Alloys
(Cohesion and Structure) vol 3, ed F R de Boer and
D G Pettifor (Amsterdam: North-HollandElsevier)
[3] Bozorth R M 1953 Ferromagnetism (New York: Van
Nostrand)
[4] Moran-Lopez J L and Falicov L M 1979 Solid State
Commun. 31 325
[5] Moran-Lopez J L and Falicov L M 1980 J. Phys. C: Solid
State Phys. 13 1715
[6] Van Hove M A, Koestner R J, Stair P C, Biberian J P,
Kesmodel L I, Bartos and Somorjai G A 1981 Surf.
Sci. 103 189
[7] Vasiliev M A 1988 Structura i Dinamika Poverkhnosti
Perekhodnykh Metallov (Kiev: Naukova Dumka) (in
Russian)
3067

M A Vasiliev
[8] Valenta L and Sukiennicki A 1966 Phys. Status Solidi 17
903
[9] Nielsen P E H 1973 Phys. Lett. 42A 468
[10] Sundaram V S, Farrell B, Alben R S and Robertson W D
1973 Phys. Rev. Lett. 31 1136
[11] Gauthier Y, Baudoing-Savois R and Bugnard J M 1992
Surf. Sci. 276 1
[12] Bardi U 1994 Rep. Prog. Phys. 57 939
[13] Hansen M and Anderko K 1958 Constitution of Binary
Alloys (New York: McGraw-Hill)
[14] Potter H C and Blakely J M 1975 J. Vac. Sci. Technol. 12
635
[15] Buck T M, Wheatley G H and Marchut L 1983 Phys. Rev.
Lett. 51 43
[16] Krummacher S, Sen N, Gudat W, Johnson R, Grey F and
Ghijsen J 1989 Z. Phys. B 75 235
[17] McRae E G and Buck T M 1990 Surf. Sci. 227 67
[18] McRae E G and Malic R A 1990 Phys. Rev. B 42 1509
[19] McRae E G, Buck T M, Malic R A and Wallace W E
1990 Surf. Sci 238 L481
[20] Shaw C G and Fain S C Jr 1977 Proc. 7th Int. Vacuum
Congress 3rd Int. Conf. on Solid Surfaces (Vienna)
ed R Dobrozemsky, F G Rudenauer, F P Viekbock and
A Breth, p 535
[21] McRae E G and Malic R A 1984 Surf. Sci. 148 551
[22] Jamison K D, Lind D M, Dunning F B and Walters G K
1985 Surf. Sci. 159 L451
[23] Alvarado S F, Campagna M, Fattah A and Velhoff W
1987 Z. Phys. B 66 103
[24] Dosch H, Mailander L, Lied A and Peisl J 1988 Phys.
Rev. Lett. 60 2382
[25] Sundaram V S and Robertson W D 1976 Surf. Sci. 55 324
[26] McRae E G, Buck T M and Malic R A 1987 Phys. Rev. B
4 2341
[27] Lipowsky R 1982 Phys. Rev. Lett. 49 1575
[28] Lipowsky R and Speth W Phys. Rev. B 28 3983
[29] Legrand B and Treglia G 1990 Phys. Rev. B 41 4422
[30] Lipowsky R 1983 Z. Phys. B 51 165
[31] Buck T M 1987 The Structure of Surfaces II (Springer
Ser. Surf. Sci. 11) (Berlin: Springer) p 435
[32] Sanchez J M and Morian-Lopez J L 1985 Surf. Sci. 157
L297
[33] Teraoka Y 1991 Surf. Sci. 256 L579
[34] Reichert H, Eng P J, Dosch H and Robinson I K 1995
Phys. Rev. Lett. 74 2006
[35] Mecke K R and Dietrich S 1995 Phys. Rev. B 52 2107
[36] Jordan R G, Sohal G S, Gyorffy B L, Durham P J,
Temmerman W M and Weinberger P 1985 J. Phys. F:
Met. Phys. 15 L135
[37] Jordan R G 1988 Vacuum 38 267
[38] Jordan R G and Sohal G S 1982 J. Phys. C: Solid State
Phys. 15 L663
[39] Wertheim G K 1987 Phys. Rev. B 36 4432
[40] Schneider U, Castro G R and Wandelt K 1993 Surf. Sci.
287/288 146
[41] Castro G R, Schneider U, Busse H, Janssens T and
Wandelt K 1992 Surf. Sci. 269/270 321
[42] Schneider U, Castro G R, Busse H, Janssens T,
Wesemann J and Wandelt K 1992 Surf. Sci. 269/270
316
[43] Shen Y G, OConnor D J, McDonald R J and Wandelt K
1995 Solid State Commun. 96 557
[44] Beccat P, Gauthier Y, Baudoing-Savois R and
Bertolini J C 1990 Surf. Sci. 238 105
[45] Baudoing-Savois R, Gauthier Y and Moritz W 1991 Phys.
Rev. B 44 12 977
[46] Hammar M, Gauthier V, Gothelid M, Karlsson U O,
Flodstrom S A and Rosengren A 1993 J. Phys.:
Condens. Matter 5 2837
[47] Moran-Lopez J L and Wise H 1980 Appl. Surf. Sci. 4 93
[48] Allie G and Lauroz C 1982 Surf. Sci. 116 L179
[49] Bevolo A I 1983 J. Vac. Sci. Technol. A 1 1004
3068

[50] Asami K and Hashimoto K 1984 Corros. Sci. 24 83


[51] Mrozek P, Menyhard M, Jablonski A and Tyuliev G 1988
Phys. Status Solidi a 110 495
[52] Wandelt K and Ertl G 1976 J. Phys. F: Met. Phys. 6 1607
[53] Vasiliev M A and Gorodetsky S D 1986 Surf. Sci. 171 543
[54] Sanchez J M and Moran-Lopez J L 1985 Phys. Rev. B 32
3534
[55] Vasiliev M A and Gorodetsky S D 1987 Vacuum 37 723
[56] Vasiliev M A and Gorodetsky S D 1996 Surf. Rev. Lett. 3
1779
[57] Baird R J 1984 Bull. Am. Phys. Soc. 29 268
[58] Feder R 1981 J. Phys. C: Solid State Phys. 14 2049
[59] Alvarado S F, Campagna M and Hopster H 1982 Phys.
Rev. Lett. 48 51
[60] Dresselhaus M S 1981 Nature 292 196
[61] Thapliyal H V and Blakely J 1978 J. Vac. Sci. Technol. 15
600
[62] Jach T and Hamilton J C 1982 Phys. Rev. B 26 3766
[63] Vasiliev M A and Gorodetsky S D 1987 Phys. Status
Solidi a 102 K121
[64] Mamaev Yu A, Petrov V N and Starovoitov S A 1987
Pis. Zh. Tekh. Fiz. 13 1530 (in Russian)
[65] Binder K, Stauffer D and Wildpaner V 1975 Acta Metall.
23 1191
[66] Schmid M, Hofer W, Varga P, Stoltze P, Jacobsen K W
and Norskov 1995 Phys. Rev. B 51 10 937
[67] Voges D, Taglauer E, Dosch H and Peisl J 1992 Surf. Sci.
269/270 1142
[68] Lui C T, Cahn R W and Sauthoff G (eds) 1992 Ordered
IntermetallicPhysical Metallurgy and Mechanical
Behaviour (NATO ASI Series E: Applied Sciences 213)
(Deventer: Kluwer)
[69] Sondericker D, Jona F, Moruzzi V L and Marcus P M
1985 Solid State Commun. 53 175
[70] Sondericker D, Jona F and Marcus P M 1986 Phys. Rev.
B 33 900
[71] Sondericker D, Jona F and Marcus P M 1986 Phys. Rev.
B 34 6775
[72] Sondericker D, Jona F and Marcus P M 1986 Phys. Rev.
B 34 6770
[73] Mullins D R and Overbury S H 1988 Surf. Sci. 199 141
[74] Davis H L and Noonan J R 1987 The Structure of Surfaces
II (Springer Ser. Surf. Sci. 11) (Berlin: Springer) p 152
[75] Miedema A R 1978 Z. Metallk. 69 455
[76] Davis H L and Noonan J R 1985 Phys. Rev. Lett. 54 566
[77] Yalisove S M and Graham W R 1987 Surf. Sci. 183 556
[78] Wuttig M 1990 J. Electron Spectrosc. and Relat. Phenom.
54/55 383
[79] Noonan J R and Davis H L 1987 Phys. Rev. Lett. 59 1714
[80] Niehus H, Raunau W, Besocke K, Spitzl R and Comsa G
1990 Surf. Sci. 225 L8
[81] Bardi U and Ross P N 1984 Surf. Sci. 146 L555
[82] Paul J, Cameron S D, Dwyer D J and Hoffmann F M
1986 Surf. Sci. 177 121
[83] Atrei A, Bardi U, Rovida G, Torrini M and Zanazzi E
1992 Phys. Rev. B 46 1649
[84] Chen W, Paul J, Barbieri A, Van Hove M A, Camerson S
and Dwyer D J 1993 J. Phys.: Condens. Matter 5 4585
[85] van Santen R A and Sachtler W M H 1974 J. Catal. 33
202
[86] Haner A N and Ross P N 1991 Surf. Sci. 249 15
[87] Atrei A, Bardi U, Torrini M, Zanazzi E, Rovida G,
Kasamura H and Kudo M 1993 J. Phys.: Condens.
Matter 5 L207
[88] Bouwman R, Toneman L H and Holscher A A 1973 Surf.
Sci. 35 8
[89] Fujinaga Y 1979 Surf. Sci. 86 581
[90] Anderson G W, Pope T D, Jensen K O, Griffiths K,
Norton P R and Schultz P J 1993 Phys. Rev. B 48
15 283
[91] Pope T D, Griffiths K, Zhdanov V P and Norton P R
1994 Phys. Rev. B 50 18 553

Surface effects
[92] van Langeveld A D, Hendrickx H A C M and
Nieuwenhuys B E 1983 Thin Solid Films 109 179
[93] Newton M A, Francis S M, Li Y, Law D and Bowker M
1991 Surf. Sci. 259 45
[94] Newton M A, Francis S M and Bowker M 1991 Surf. Sci.
259 56
[95] Holmes D J, King D A and Barnes C J 1990 Surf. Sci.
227 179
[96] Lindroos M, Barnes C J, Bowker M and King D A 1991
The Structure of Surfaces III (Springer Ser. Surf. Sci.
24) ed S Y Tong et al (Berlin: Springer) p 287
[97] Rao R S, Bansil A, Asonen H and Pessa M 1984 Phys.
Rev. B 29 1713
[98] Gonis A 1986 Phys. Rev. B 34 8313
[99] Zhang X, Hwang M, Gonis A and Freeman A J 1986
Phys. Rev. B 34 5169
[100] Wright H, Weightman P, Andrews P T, Folkerts W,
Flipse C F J, Sawatzky G A, Norman D and
Padmore H 1987 Phys. Rev. B 35 519
[101] Weightman P, Wright H, Waddington S D,
van der Marel D, Sawatzky G A, Diakun G P and
Norman D 1987 Phys. Rev. B 36 9098
[102] Gonis A, Zhang X-G, Freeman A J, Turchi P, Stocks G M
and Nicholson D M 1987 Phys. Rev. B 36 4630
[103] Zhang X-G and Gonis A 1989 Phys. Rev. Lett. 62 1161
[104] Sowa E C, Gonis A, Zhang X-G and Foiles S M 1989
Phys. Rev. B 40 9993
[105] Margi R, Wei S-H and Zunger A 1990 Phys. Rev. B 42
11 388
[106] Gonis A, Zhang X-G, MacLaren J M and Czampin S
1990 Phys. Rev. B 42 3798
[107] Lu Z W, Wei S-H, Zunger A, Frota-Pessoa S and
Ferreira L G 1991 Phys. Rev. B 44 512
[108] Lu Z W, Wei S-H and Zunger A 1991 Phys. Rev. B 44
3387
[109] Lu Z W, Wei S-H, Zunger A and Ferreira L G 1991 Solid
State Commun. 78 583
[110] Gonis A 1992 Green Functions for Ordered and
Disordered Systems (Amsterdam:
North-HollandElsevier)
[111] Lu Z W, Wei S-H and Zunger A 1992 Phys. Rev. B 45
10 314
[112] Johnson D D and Pinski F J 1993 Phys. Rev. B 48 11 553
[113] Barbier L, Salanon B and Loiseau A 1994 Phys. Rev. B
50 4929
[114] Bowker M, Newton M, Francis S M, Gleeson M and
Barnes C 1994 Surf. Rev. Lett. 1 569
[115] Weightman P, Cole R J, Verdozzi C and Durham P J 1994
Phys. Rev. Lett. 72 793
[116] Clark J F, Pinski F J, Johnson D D, Sterne P A,
Staunton J B and Ginatempo B 1995 Phys. Rev. Lett.
74 3225
[117] Gonis A, Singh P P, Turchi P E A and Zhang X-G 1995
Phys. Rev. B 51 2122
[118] Loboda-Cackovic J 1995 Vacuum 46 89
[119] Loboda-Cackovic J 1995 Vacuum 46 1449
[120] Weightman P, Cole R J, Brooks N J and Thorntor J M C
1995 Nucl. Instrum. Methods Phys. Res. B 97 472
[121] Cole R J, Brooks N J, Weightman P, Francis S M and
Bowker M 1996 Surf. Rev. Lett. 3 1763
[122] Gallis C, Legrand B, Treglia G, Salanon B, Hecquet P
and Saul A 1996 Surf. Sci. 352 588
[123] Loboda-Cackovic J 1996 Vacuum 47 1405
[124] Bardi U, Atrei A, Ross P N, Zanazzi E and Rovida G
1989 Surf. Sci. 211/212 441
[125] Bardi U, Beard B C and Ross P N 1988 J. Vac. Sci.
Technol. A 6 665
[126] Bugnard J M, Baudoing-Savois R, Gauthier Y and
Hill E K 1993 Surf. Sci. 281 62
[127] Gauthier Y, Hoffmann W and Wuttig M 1990 Surf. Sci.
233 239
[128] Gauthier Y, Baudoing R, Lundberg M and Rundgren J

1987 Phys. Rev. B 35 7867


[129] Gauthier Y, Baudoing R and Jupille J 1989 Phys. Rev. B
40 1500
[130] Gauthier Y, Joly Y, Baudoing R and Rundgren J 1985
Phys. Rev. B 31 6216
[131] Baudoing R, Gauthier Y, Lundberg M and Rundgren J
1986 J. Phys. C: Solid State Phys. 19 2825
[132] Gauthier Y and Baudoing R 1990 Segregation and
Related Phenomena ed P Dowben and A Miller (Boca
Raton, FL: Chemical Rubber Company) p 169
[133] De Temmerman L, Creemers C, Van Hove H, Neyens A,
Bertolini J C and Massardier J 1986 Surf. Sci. 178 888
[134] van de Riet E, Deckers S, Habraken F H P M and
Niehaus A 1991 Surf. Sci. 243 49
[135] Jugnet Y, Prakash N S, Grenet G and Duc T M 1987
Surf. Sci. 189/190 649
[136] Schmid M, Stadler H and Varga P 1993 Phys. Rev. Lett.
70 1441
[137] Gauthier Y, Baudoing R, Joly Y, Rundgren R,
Bertolini J C and Massardier J 1985 Surf. Sci. 162 342
[138] Bertolini J C et al 1982 Surf. Sci. 119 95
[139] Ferrante J 1971 Scr. Metall. 5 1129
[140] Ferrante J 1971 Acta Metall. 19 743
[141] Pessa M, Asonen H, Rao R S, Prasad R and Bansil A
1982 Surf. Sci. 117 371
[142] Baird R J, Ogletree D F, Van Hove M A and
Somorjai G A 1986 Surf. Sci. 165 345
[143] Esposto F J, Zhang C-S and Norton P R 1993 Surf. Sci.
290 93
[144] Sukiennicki A 1973 Phys. Lett. 44A 169
[145] Sukiennicki A and Puszkarski H 1969 Acta Phys. Pol. 36
581
[146] Vasiliev M A, Gorodetsky S D and Blaschuk A G 1993
Kristallografiya 38 205 (in Russian)
[147] Kumar V and Bennemann K H 1984 Phys. Rev. Lett. 53
278
[148] Moran-Lopez J L and Bennemann K H 1977 Phys. Rev. B
15 4769
[149] Balseiro C A and Moran-Lopez J L 1980 Phys. Rev. B 21
349
[150] Mezey L Z, Nishikawa O and Giber J 1983 Phys. Status
Solidi a 78 323
[151] Kroll D M and Gompper G 1987 Phys. Rev. B 36 7078
[152] Teraoka Y 1987 Solid State Commun. 64 1333
[153] Teraoka Y 1990 Surf. Sci. 233 97
[154] Teraoka Y 1990 Surf. Sci. 232 193
[155] Rivers S B, Unertl W N, Hung H H and Liang K S 1995
Phys. Rev. B 52 12 601
[156] Hillert M and Rundgren J 1985 Phys. Rev. B 32 640
[157] Legrand B and Treglia G 1987 The Structure of Surfaces
II (Springer Ser. Surf. Sci. 11) (Berlin: Springer) p 434
[158] De Temmerman L, Creemers C, Van Hove H and
Neyens A 1987 Surf. Sci. 183 565
[159] Lundberg M 1987 Phys. Rev. B 36 4692
[160] Eymery J and Joud J C 1990 Surf. Sci. 231 419
[161] Mezey L Z and Hofer W 1992 Surf. Sci. 269/270 1135
[162] Stadler H, Hofer W, Schmid M and Varga P 1993 Surf.
Sci. 287/288 366
[163] Stadler H, Hofer W, Schmid M and Varga P 1993 Phys.
Rev. B 48 11 352
[164] Abraham F F, Tsai N-H and Pound G M 1979 Surf. Sci.
83 406
[165] Wynblatt P and Ku R C 1977 Surf. Sci. 65 511
[166] Treglia G and Legrand B 1987 Phys. Rev. B 35 4338
[167] Walker R A and Darby J B 1970 Acta Metall. 18 1261
[168] Spencer M S 1984 Surf. Sci. 145 145
[169] Williams F L and Nason D 1974 Surf. Sci. 45 377
[170] Teraoka Y 1990 Surf. Sci. 235 249
[171] Huberman B A and Streifer W 1975 Phys. Rev. B 12 2741
[172] Tahir-Kheli R A and Kawasaki T 1977 J. Phys. C: Solid
State Phys. 10 2207
[173] Binder K, Bowker M, Inglesfield J F and Rous P J 1995
3069

M A Vasiliev

[174]
[175]
[176]
[177]
[178]
[179]
[180]
[181]
[182]
[183]
[184]
[185]

3070

Cohesion and Structure vol 4, ed F R de Boer and


D G Pettifor (Amsterdam: North-HollandElsevier)
Urias J and Moran-Lopez J L 1982 Phys. Rev. B 26 2669
Mathon J 1988 Rep. Prog. Phys. 51 1
Gradmann U 1991 J. Magn. Magn. Mater. 100 481
Gradmann U 1993 Handbook of Magnetic Materials vol 7,
ed K H Buschow (Amsterdam: Elsevier) pp 196
Sukiennicki and Wojtczak L 1973 Phys. Rev. B 7 2205
Binder K and Hohenberg P C 1974 Phys. Rev. B 9 2194
Binder K, Landau D P and Kroll D M 1986 Phys. Rev.
Lett. 56 2272
Binder K and Hohenberg P C 1972 Phys. Rev. B 6 3461
Chame A and Tsallis C 1989 J. Phys.: Condens. Matter 1
10 129
Auilera-Ganja F and Moran-Lopez J L 1985 Phys. Rev. B
31 7146
Landau D P and Binder K 1990 Phys. Rev. B 41 4633
Moran-Lopez J L and Falicov L M 1979 Phys. Rev. B 20
3900

[186] Moran-Lopez J L and Falicov L M 1980 J. Magn. Magn.


Mater. 1518 1077
[187] Urias J and Moran-Lopez J L 1991 Phys. Rev. B 23 2759
[188] Moran-Lopez J L and Falicov L M 1979 Phys. Rev. B 19
1470
[189] Kaneyoshi T, Tamura I and Sarmento E F 1983 Phys.
Rev. B 28 6491
[190] Moran-Lopez J L and Falicov L M 1978 Phys. Rev. B 18
2542
[191] Gradmann U, Waller G, Feder R and Tamura E 1983
J. Magn. Magn. Mater. 3134 883
[192] Victora R H 1986 Magnetic Properties of
Low-Dimensional Systems ed L M Falicov and
J L Moran-Lopez (Berlin: Springer) p 25
[193] Mailander L, Dosch H and Peisl J 1990 Phys. Rev. Lett.
64 2527
[194] Yurasova V 1983 Vacuum 33 565
[195] Kuvakin M, Yurasova V and Motaweh H 1994 Vacuum
45 45

Anda mungkin juga menyukai