You are on page 1of 66

Chapter 1

STRUCTURE AND PROPERTI ES OF METALS AND ALLOYS


1.1 A microscopic theory of solids
1.1.1 The quantum theory of pure metals
It is impossible to present in a single chapter an exact theory of metals and alloys,
or of phenomena, such as those forming the basis of electron-spectroscopies, that are used
to study and to establish the electronic structure of metallic catalysts. However, it is felt
that a book on catalysis by alloys should at least introduce some of the important terms
(band structure, density of states, photoemission from the valence band, etc.) on basis of
some very simple theoretical considerations; it is not our ambition to achieve more than
that.
All modern books on undergraduate physical chemistry [1-3] offer an introduction
to quantum mechanics, which is the basis of the chemical bond theory. The reader is thus
expected to be familiar with terms such as the wave or state function (e.g. Z or ~t), the
Hamiltonian or total energy operator It/and the Schr6dinger equation:
(/-~ - E) z =0 (1)
where E is the steady state total energy of the system, the state of which is described by
function Z. The total energy operator 121 can be split into two parts, the kinetic energy
operator T and the potential energy operator V. Operator T is a differential operator and
thus equation 1 is a differential equation of second order. The text books [1-3] offer also
an introduction to the form of functions Z for hydrogen atom, for the hydrogen-like atoms
(lithium, sodium, potassium, etc.) and for functions (orbitals) of some other atoms. With
metals and alloys we are interested in the form of the solid or crystal orbitals, ~. Let us
summarize some of their basic features [4].
We shall mostly be interested in metals or alloys in a crystalline form. Such bodies
distinguish themselves by having a periodic potential V, so if we consider a linear one -
dimensional system of an infinite length and with a periodicity, i.e. lattice constant, a, the
electron density p, which is proportional to the probability of finding an electron in a unit
volume, i.e. gt*gt where gt* is the complex conjugated form of ~, will be the same at all
places differing by a; therefore we state that
8 chapt er 1
p = ~*(x +a)~(x+a) = ~*(x)~(x)
(2)
This means that ~( x+a) and ~( x) differ in such a way that
~( x+a) = X, ~( x+2a) = xz . . . . . (3)
,(x) ,(x)
and ~,*~, equals 1. The factor ~, is either unity, whi ch is a trivial case, or it is a compl ex
number, whi ch is the general case. If we assume, fol l owi ng the idea of Born and Karman,
that in a crystalline solid when passing a region of N atoms, then we find a point where
not only the densities but also the wave functions t hemsel ves are equal, that is ~( x+Na) =
~(a), then the most general form of ~,, which is the one we are seeking, is
~= exp (i 2n g/N) = exp (i 2r~ galL)
(4)
where g is 0,1,2... N-1 and can be regarded as the quant um number descri bi ng the
different ~' s. The symbol L stands for Na. Thus
X= ~(x+a) = exp ( i 2z ga) . (5)
,(x) L
We are not interested in the ratio X, but in the most general form of ~(x). Then we have
to realize that this must contain a factor like the exponent i al t erm in equat i on 5 and there
may be a function U(x) havi ng the property that U(x+a) = U(x). A funct i on with such
propert y woul d cancel in the ratio ~, in equation 5. Wi t h a very large value of a, U(x) is
the at omi c orbital. The Bl och t heorem states that the most general form of the funct i on
~t(x) is
, ( x) = oxp (th). U(x)
(6)
and, i ndeed this form, called the Bl och function, fulfils all the conditions whi ch we have
put on ~( x) above. Vari ous trial forms of equation 6 can be used and various degrees of
approxi mat i on accordi ngl y achieved. When U(x) is put equal to a constant, ~( x) = A exp
(ikx), and for three di mensi ons we write: ~( r ) = A exp(ikr).
The funct i on ~( r ) is a plane wave and this is the free electron approximation. When U(x)
is a series of exponentials, i.e. a Fourier-like series, we have the nearly-free electron
approximation. Some other hi gher approxi mat i ons will be listed bel ow, but first we shall
turn our attention to the ' one electron approxi mat i on' which chemi st s know well as the
linear combi nat i on of atomic orbitals (LCAO) and we shall apply it to our one-di mensi -
onal solid chain of atoms.
Structure and properties of metals and alloys 9
We substitute the crystal orbital by a linear combination of atomic orbitals I~ n
n indicating their place
with
Vrr = L.C.A.O. = n Cn *n (7)
in the Schr6dinger equation 1 and read it as:
2 n C n ( f l - E) On -" 0 (8)
To convert equation 8, which is a differential equation, into an algebraic relation bet ween
numerical values, we multiply it by 0*e and integrate over the space of N atoms. In this
way, by writing t7t in an extended form (the usefulness of it will be seen immediately), we
obtain the relation
]~n Cn {I **e (T "1- Vcrystal + Uat- Uat ) 0Pnd~ - g I **e *n d'l:} = 0
(9)
T + Uat is H ~ the Hamiltonian of a free atom of the element of the chain and d'c the
element of the space. Substituting this expression for H ~ and calling Vcryst- Uat "- AV, we
obtain
]~n Cn {I 0*e ( H~ AV) *n d'l~- g I 0*e On d'c} = 0
(10)
We now assume that the overlap between *n and **e is zero when n is not equal to e (n
and e denote different positions of atoms), but is unity when e equals n. Thus
I 0*e *n d'c = o and I **n *n d'c = 1 (11)
Further, in our approximation we keep only the following terms of equation 10
-- Eat (e=n) =
II~e H~ Cn d1: (12)
= 0 (ee:n)
= 13 (e=n_+l)
I *e mv *n = 0 (e=n) (13)
= 0 (e>n_+l)
10 chapter 1
The second term (e=n) is zero because of the definition of the ' zero' potential energy. If
we choose n, somewhere in the chain, we are then left with a general equation for Cn:
(14)
since all other terms vanish in our approximation defined by equations 11 to 13. Then
a n + 1 + a n _ 1
E= a + [3 (15)
Cn
The Bloch theorem can be fulfilled by taking
C,, = exp (i kx) (16)
since
~kx,, i kx -x)
0 = ~ e gO,,=e . ~ e rex" gO(x-x,,) (17)
n n
The term represented by the summation has the properties of the function U(x). By
substituting the expression for Cn in equation 15 the energy E is
E(k)=a + [3(e-~+e +U'a) =a +213 coska
(18)
Several features of this equation are very interesting. There are N different values of k,
thus there are N different crystal orbitals ~k and N different energy levels. The levels E(k)
form, for a large Ns, a quasi-continuous band of energies between Ema x and Emin, and from
equation 18 the band width is
Emax-Em~n=4p (19)
In other words, the band width is proportional to the overlap or hopping integral B. The
(n-1)d orbitals (n being the principal quantum number of valence electrons) overlap less
than do the ns orbitals. In the rough approximation which we have considered, the bands
are separated and the (n-l) d-band is narrow and the (n) s-band is broad. The d-band has
then 5N levels, but the s-band only N levels. The density of states between E and E+dE is
as a consequence higher in the d-band than in the s-band. These are the main pieces of
information which one needs in order to understand chapters 2 and 3.
Structure and properties of metals and alloys 11
E(k)
l
- l ' t 0
O
k !
O
>N(E)
>N( E)
f i gure 1
left: Energy as a f unct i on of the wave-vector k, f or a hypothetical one-dimensional
' crystal' (a chain of atoms) with a lattice constant a.
right: Densi t y of states curve, corresponding to E(k) shown in the left part.
The function cos (k a) is defined over the interval k from -rt/a to +rt/a and in figure
1 6 has a negative value, and tx is taken as the arbitrary zero. Where the slope of the
function is steep, t here are only a few states (i.e. few k' s) in a given range of energy, but
where the slope is low, as in the neighbourhood of g/a, there are many. In other words,
the density of states N(E) is a function which increases with (dE/dk) -1, as is seen in figure
2. We shall meet the term density of states at many places in this book.
In the free electron approximation and for an one dimensional solid chain, ~k is
equal to A e ikx and the Schr6dinger equation is used with V equal to zero. It reads
d21~k 8n2m
+ ~ E q ~ k = 0
d~ 2 h 2
(20)
Substitution of ~k by the function of free electrons in the Schr6dinger equation produces
12 chapter 1
kZh 2
E - (21)
8~;2m
which can be compared with the Newtonian relation which says that E equals pZ/2m, with
p being the momentum. Indeed, for free electrons the momentum p equals h k or in other
words, k is a moment um in units of h.
El k)
- n k 0 k n
G t3
a 2 - 2r~ 1
a 2
a 2 + 2132
a l - 2 ~ 1
a l + 2 [31
f i gur e 2
E(k), di spersi on f unct i ons, f o r a
hypot het i cal one di mens i onal
cryst al ( at omi c chai n) wi t h t wo
orbitals, correspondi ng to the
energi es (z 1 and (z 2 on each
atom.
j(31, j(32 < O; I~11< Ij% l
The interval -n/ a to n/ a forms a k-space into which all possible non-equivalent k' s
are placed; it is called the first Brillouin zone. We have seen through our discussion of the
Bloch theorem that k equals 2ng/Na, or in other words k is related to the reciprocal lattice
constant, a -1. The interval -n/ a to n/ a is thus reciprocal with respect to real space. Because
p equals hk, this interval is also a momentum space, into which all possible states, each
characterized by its energy and k-number or momentum, are placed.
In higher approximations than that corresponding to the free electron model, the
moment um is not equal to hk. However, with regard to various forces F, hk still behaves
like a momentum, since F is always equal to hk' . Therefore, k can be called a pseudo-mo-
mentum, in units of h. This is an important point for understanding the angle-resolved
valence-band photoemission, which is discussed in chapters 2 and 3.
Let us now make the step from chains of atoms to two-dimensional flat arrays.
Now, for a square lattice
Structure and properties of metals and alloys 13
~ , = P ' C r ,~ ~,.,~ (22)
r, s
with
O r , s = e Xp ( i k r X r + i k s Y s ) (23)
and the energy is
E( k ) = a + 213(cosk~a + coskya)
(24)
and it forms a plane in the E(kx, ky) space (see figure 3, under a) on the left)
The first Brillouin zone shown under b) in fig.3 is now two dimensional, as is also
the whole k-space. It is often useful to show the function E(k) in a more simple way.
Then E(k) is calculated for selected values of k, for example along certain lines in the
Brillouin zone, such as, from the point k(0,0) to the point k0t/a; 0), etc., as is shown in
figure 3 left under c). The density of states corresponding to the energies between ~ + 413
and ot - 4g is shown in the lower right comer [4].
In three dimensions, with E (k x, ky, kz) the pictures are slightly more complicated
but not conceptually very different.
The mathematical theory of groups teaches us that in each space having translati-
onal symmetry and in which Born-Karman conditions hold, there are always 14 different
Bravais lattices. Both the real and the reciprocal or k space are spaces with a lattice, i.e.
translational symmetry, and this means that each lattice type in one space must have a
counterpart in the other (i.e. reciprocal) space. For example, it follows that bcc (real space)
translates into fcc (reciprocal) and fcc (real) translates into bcc (reciprocal), etc. Further,
it follows that the vector k and the pseudo-momentum p have in a crystal of rectangular
form the same directions in the real and in the reciprocal space. This enables us to
indicate the real movement of electrons by movements of k-states in the reciprocal space.
This is again an important statement for the description and understanding of the angle-
resolved electron photoemission. If electrons proceed in the k-space derived, for example,
from a cubic crystal in a certain direction, they do the same in the real crystal too, and at
the surface they continue to pass into vacuum without refraction because the kx,ky-compo-
nents are preserved. Figure 4 shows a Brillouin zone of an fcc real-space lattice.
The function E(k), called often the dispersion law, is usually theoretically calcula-
ted for certain selected values of k, for example, for k' s along well chosen lines intercon-
necting important points on the Brillouin zones. These points are denoted by letters F, X,
K, W etc. ( see also chapters 2 and 3). The typical form of such E(k)-sections are shown
in figure 5a by results for copper. When the k-points corresponding to the highest energy
levels still occupied by electrons are interconnected by a plane, the so called Fermi surface
14 chapter 1
is created (figure 5b).
(a)
k,
- - t~/ a
k~
' a
ky
(b)
bl[ / a
- - : ~/ a
~/ a w kl
Energy
(c)
=-4~
o{
~+4~
(0,0) (n,a, O) (n/a,n/a) (0,0)
X M F
NIE)
f i gure 3.
A hypothetical two dimensional crystal.
Three representations of E(k) f or the s band.
(a) Energy surface f or one quarter of the Brillouin zone.
(b) Constant-energy contours, illustrating the symmetry of the zone.
(c) Energy pl ot t ed over a triangular path of k values, showing mi ni mum and maxi mum
energies, and density of states.
(for symbols F, X and M see f i gure 4)
Structure and properties of metals and alloys 15
k
z
1
t d l
ky
figure 4.
Brillouin zone in a reciprocal space with b.c.c.
lattice, corresponding to f c. c. lattice in the
real space.
figure 5.
Band structure (a) and
Fermi surface (b) f or copper.
In (a) the Cu 3d bands are
labelled; the dashed curve
shows the 4s band predicted
without any mixing with
the d band [4]
E nergy [
F X W
J
C " 3 ' ' '
,,,,~
I" K
{b)
16 chapter 1
The occupation of E(k)-levels at temperatures above absolute zero is governed by
the Fermi-Dirac distribution function [1-4].
E - E v ) ] (25)
f ( E ) = [ 1 + e x p ( -1
k T
Equation 25 states that as the temperature approaches zero, all levels below EF, the Fermi
energy, become occupied (i.e. f(E) tends to unity) and all levels above E v become vacant.
Thus for metals having a pseudo-continuous band, E F is the highest occupied level at the
absolute zero. By using equation 25 in statistical thermodynamics, one can derive that E F
is the total free energy of the electrons, per electron, i.e. it is the electrons' chemical
potential. At equilibrium there is only a single value of EF for the whole system.
The Fermi energy is a total energy, i.e. it includes also the electrostatic potential
energy, such as that due to the contact potential between metals, and other similar terms.
It is therefore almost always necessary to take EF as the zero reference level, because it is
always difficult and usually impossible to establish the exact position of EF with regard to
the vacuum level Eva c. The work function of the metal, O, is only approximately (0-2 eV)
equal to Evac-E F.
If one connects by a continuous surface in k-space all k' s corresponding to EF, the
so-called Fermi surface arises. For free electrons, as can be seen from equation 21, this
surface is a sphere (EF - k2). For other, higher approximations this sphere is deformed; an
example of a Fermi surface which has been established experimentally as well as by
theoretical calculations is shown in figure 5b.
The geometric form of the Fermi surfaces is already known for most metals [7].
The main techniques to establish the form of the Fermi surface are those associated with
the so-called de Haas and van Alphen effect and the skin effect [7].
An important feature of a metal or an alloy is the density of states at the Fermi
level N(Ev); this value can be determined by measurements of the magnetic susceptibility
and the heat capacity at low temperature [7].
In the discussion on the electronic structure of metals and alloys and its relation to
electron spectroscopies, the most important quantity is most probably the density of states,
N(E). This is because a simple relation exists between the distribution of the photoemitted
electrons I(E), and the integral density of states taken over all angles N(E): to a good
approximation
I ( E ) = c o n s t . M~i . N( E) ininarN(E) f ~ 2
(26)
where Mf, i is ~ffinal H' ~initial dx and H' is the perturbation causing the transition from the
initial into the final state. The density of states for the free electron in the final state is
proportional to ~/E, so that for high energies it changes comparatively little over the
Structure and properties of metals and alloys 17
energy range of interest. This means that the distribution I(E) measured at the detector is
only a slightly deformed density of states for the system before ionisation, N(E)initiav
In the approximation of ' nearly free electrons' , the density of states function
resembles that shown in figure 6. This simple form already reflects the main features
observed experimentally, and therefore in theoretical discussions the N(E) function is often
schematically pictured as in figure 6 (see chapters 2 and 3).
Figure 5b shows the Fermi surface of copper. If this were a metal which could be
exactly described by a free electron model, the Fermi surface would be perfectly spherical.
The " necks" sticking out towards the (111) faces are caused by the periodic crystal
potential V.
figure 6.
Density of states in a band
(Ema ~ Emin) f or a model of
'nearly f ree' electrons
NI E)
i
I .
s
i \
/ I
\
I
I
I
I , / E
I
, , J
E mi n E ma x
- E
ma x
E
The volume of the space enclosed by the Fermi surface depends on the total
number of electrons n in the system; for free electrons, the surface area is proportional to
n ~'3. Knowing that, we shall now make a hypothetical experiment: we replace some copper
atoms by atoms with more than one valence electron, for example, by zinc or aluminum.
This increases the volume under the E F surface and since the sphere cannot in general case
continue to grow into the higher Brillouin zone, because of a gap in energy on the
Brillouin zone face, the sphere-like form will probably be deformed to fill up the states
near to and just under the Brillouin zone faces. However, it is also possible that if the
alloy could have another crystallographic structure than that of copper, and as a conse-
quence to have another form of Brillouin zone, the additional electrons could be better
accommodated at lower energies, for example, in a more sphere-like body. Thus, the
average number of electrons per atom in such cases will dictate the crystallographic
structure of the alloy. Hume-Rothery has formulated several very useful rules relating the
most stable structure to the average number of electrons [8], and although some details of
his theory are not longer valid, the basic idea is probably sound. There are also some
papers which try to relate the Hume-Rothery' s structural changes in alloys to the changes
18 chapter 1
in the catalytic activity [9].
Three approximations in the description of the behaviour of electrons in a periodic
potential have so far been mentioned: (i) a model of free electrons; (ii) a model of nearly
free electrons; (iii) a model of electrons tightly bound to the atoms (tight binding
approximation, with L.C.A.O. used as a trial function).
These approximations are useful to elucidate the terms of which the theory and an
experimentalist make use and to identify the phenomena typical for systems with a
periodic potential, but all three approximations mentioned are unsuitable for quantitative
predictions. Approximations i and ii exaggerate the de-localization of electrons, while
approximation iii considers the electrons as too strongly bound and too much localized on
the individual atoms. Since all three are one-electron approximations and take the electron-
electron interactions (Coulombic and exchange interactions) implicitly in the average
potential, they do not treat this particular aspect properly. The higher approximations try to
improve on this situation. Of many ways of doing it that are described in the literature, we
shall mention only the following ones [10-16]: (1) Augmented Plane Wave (APW) and
related theories, [10,1] and the Korringa-Kohn-Rostoker (KKR) approximations [12],
which both attempt to improve the construction of the wave function; (2) electron density
method (Kohn, Sham, Lang [13,14]) which explicitly treats the electron-electron interacti-
ons. Just a few remarks about these theories follow.
One possibility for improving the constructed wave function is to cut the crystal in
a space where the electrons behave as essentially free and in a space where they behave as
being bound tightly to the nuclei: this is what the APW (Augmented Plane Wave) theory
does. The Schr6dinger equation is then solved inside a spherical potential wall of a radius
R. The electrostatic potential of the nucleus is hypothetically contained in this sphere,
being zero outside. The solution for the sphere resembles that for free atoms, being a
linear combination of products of the radial functions and spherical harmonics. The
coefficients of the linear combination are then chosen in such a way that the solutions
match smoothly, on the surface of the sphere, the plane waves which describe the
behaviour of electrons outside the sphere [ 10, 11 ].
Another technique for constructing a wave function or crystal orbital which would
describe properly the delocalized character of the electrons in the metal is the theory
suggested by Korringa and by Kohn and Rostoker (KKR) [12]. In the KKR theory the
atomic spheres are again considered. We can then imagine that at the surface of a certain
atomic sphere there is a solution which we shall call the outgoing function ~out, and the
same holds for all other atomic spheres. As a consequence at the surface of our first
chosen sphere a combination CI)in of all other waves exists. The two functions CI)in and Oout
are put equal on the surface of the atomic sphere, and they are mutually related as
scattered and incident waves, with scattering depending on the potential inside the atomic
spheres.
Structure and properties of metals and alloys 19
Both the APW and the KKR techniques describe the potential inside the spheres as
an artificial potential, which has no components outside the sphere; it is called muffin-tin
potential. Both techniques have also been applied to alloys (see below).
Another successful approach to the problems of the description of the solid state
has been suggested [13-16]. The authors of these papers have shown that the system of
many electrons can be totally described by the electron density n(r), and have introduced a
function E(n(r)), by the following equation [13,14]
E ( n ( r ) ) = T [n(r)] -
(27)
N n ( r ) e 2 n ( r ) n ( r / / )
E Z e 2 f Ir_RMI d r + - - f ir_--~i i
M---1 2
+ E,o._,o" + E. xch ( n ( r ) )
In this equation T stands for the kinetic energy of the non-interacting electrons with
density n(r), the second term is for nuclei-electron interactions with the nuclei at the
positions RM in the lattice, the third term is the mutual Coulombic interaction of electrons,
Eion_io n is for Coulombic repulsion of nuclei and the last term is the exchange energy. The
function E(n(r)) has a minimum when n(r) corresponds to the ground state density and the
minimal energy is then taken as the ground state energy of the system. The practical
approximation is to write down the equation for one electron functions with an effective
potential and with the exchange term written in the so called local-density approximation.
The simplest form of the whole theory is formulated for a model with an uniform
continuous positive background, with electrons as discrete charges on it. This is the so
called Jellium model. Many problems of chemisorption and promoter effects have been
successfully attacked by this theory and many important conclusions derived [15,16]. To
our knowledge it has not been used for alloys, for which it is not well suited.
1.1.2 Pauling' s theory of pure metals
This theory was formulated [17,18] at a time when the chemical bonding was
usually described in terms of electron pairs and resonance structures, with the real
structure somewhere in between them. Physicists never responded to Pauling' s idea' s with
much enthusiasm, but all his ideas, including the theory of metals, are extremely popular
among chemists. That is the main reason why they are presented and analyzed below. The
other reason is to demonstrate that in reality it is hardly possible to avoid a more difficult
theory [4-16] by accepting one such as that of Pauling. It is not possible to use semi-
empirical approaches based on vague reasoning and yet be able to make r e l i a b l e predicti-
ons.
20 chapter 1
Pauling analyzed the crystallographic structures and distances between atoms for
various metallic elements. In order to be able to compare structures with various coordina-
tion numbers, CN (CN is 8 for bcc, 12 for fcc) and with various numbers of electrons
available for bonding (that is the valency, v). Pauling introduced the so called single bond
radius, R(1) for all elements, defined as
R(n) = R(1) - 0.600 In n (28)
where R is in A and n is the bond order. For metals the latter is
n = (v/ f. N. ) (29)
The analytical form of equation 28 and the value of 0.6,h, for the prelogarithmic constant
were derived from R(1), R(2) and R(3) of ethane, ethene and ethyne. With molecules used
for this calibration the order n was thus always greater than one, but for metals it is
always less than one. However, Pauling assumed the same equation to hold for both cases,
the constant being only slightly adjusted, from 0.7A for carbon-carbon bonds to 0.6/k for
all other bonds. The system of single bond-radii is shown in figure 7 [ 17].
Z . 5
2. 0
1.5
1.0
0 . 5
0. 0
T h e f i r s t 1 0n g p e r i o d
. o
c o ~
- \ -" -%
S o ~ ~o =
T ,-,o _ _ o =o =g ~ . . o, -~-~. o
V " ~ , - O- o _ o _ o _ O
C.,r J Fe I NiCU I I ASs;" ~r
- Mn Co Zn Ge
T h e s e c o n d lo n g p e r i o d
o
z ' \ o . _ -o '-o =_~o " ~ I ~" . . .
Nb , ~17 6 Aq J J 5 Oj n
I Tr ! Rh I Cd Sn i e I
Mo ~ Pd
Ru
o S i ng l e - b o nd me t a l l i c r a d i i
o O c t a h e d r a l r a d i i
A T e t r a h e d r a l r a d i i
t t I , 1 1 1 t
18 2 0 3 0 3 6 4 0 5 0 5 4
A Kr Xe
figure 7
Single bond-radii as calculated by Pauling for the indicated metals and semiconducting
elements [17].
Structure and properties of metals and alloys 21
In the same figure the tetrahedral radii are also plotted; they are real for s,p-
elements and fictitious for the transition metals. The straight line of the first period is
described by: R(1) = Rl(SP 3) = 1.825 - 0.043 z where z is the number of electrons
outside the argon shell. Transition metals show a contraction in R(1) and according to
Pauling [17,18] this is due to the participation of d-orbitals in the metallic bond. He
therefore introduced the concept of d-character 8(in % ), a quantity expressing exactly how
much of the bonding is due to the d-electrons. With this 8, he wrote the empirical
equation for single bond radius R(1):
R(1) = Rl ( 5, z) = 1.825-0.043z - (1.600-0.100z) 8
(30)
The form of equation 30 was chosen to describe the results and to fit the curves of R(1)
vs z shown above only for the first row of transition metals. To understand how Pauling
obtained the points necessary to derive the absolute values of the constants in equation 30
we must look to his treatment of the electronic structure of the magnetic elements iron,
cobalt and nickel.
Pauling speculated that each metal has three types of orbital: (i) atomic orbitals
into which unpaired as well as paired electrons can be placed; (ii) valence orbitals into
which electrons which form the metallic bonds are placed; (iii) metallic orbitals which are
unoccupied and which mediate " unhindered resonance" . To be able to explain the use of
fractional numbers of electrons when describing the bonding, Pauling assumed that a metal
can have several imaginary extreme structures, which are mixed in certain proportions to
give the real structure. The real structure is that which results in the experimentally-found
magnetic moments. This is illustrated by table 1 [17]. It is assumed that nickel has two
structures which are mixed in the proportions 30% magnetic nickel and 70% non-
magnetic, in which all electrons are paired. This mixture leads to the experimentally found
magnetic moment per atom of 0.6 Bohr magnetons, and in a similar way the structures of
cobalt and iron are mixed to produce the experimentally determined value of the magnetic
moment per atom.
By constructing such hypothetical structures and mixing them in the indicated way,
Pauling also calculated the 8% character, (the last column of table 1). He calculated values
for the d-character for iron, cobalt and nickel and with them he created equation 30 for
R(1). This equation has to fit all points which have been calculated from R(n)' s of
individual metals. Then, he used the R(1) values to calculate 8 for non-magnetic metals
and produced the table of valencies and % d-band character (see table 2), values of which
soon became very popular amongst chemists. There have also been attempts [21] to apply
the 8 values to explain results on alloys and even on sulphides.
22 chapter 1
table 1
Percentage d-character of cobalt, nickel and copper (Pauling theory)
(brackets indicate bonding orbitals)
Metal
Co(B)
Co(B)
Ni(A)
Ni(B)
Cu(A)
Cu(B)
Outer electrons
3d
4s I 4p
~ T T ~
T, T I / ~
T$ T T' ~- ~~. 9
T~ T~ i - - - o
~ ].
T~ T~T~ I 1 - ~
P~eso-
nance
ratio
35
65
30
70
25
75
Percentage d-character
35~oo X z~ + 6~o o X 3/~ = 39.5%
30~00 X 2/~ -Jl- 70~00 X 3/~ _-400-/0
25/~00 X 3/~ _Jr- 75/~00 X 2/~ -- 35. 7%
table 2
Percentage d-character (d% ) and valency (v) of elements in the first series of transition
metals
V d%
Sc 3 20
Ti 4 27
V 5 35
Cr 6.3 39
Mn 6.4 40.1
Fe 5.78 39.7
Co 6 39.5
Ni 6 40.0
Cu 5.5 36
However, the question is whether the popularity of the 5 values is justified. They
have been derived from hypothetical electronic structures using empirical equations for
Structure and properties of metals and alloys 23
R(1)' s. It is doubtful whether equation 28 from which the argument starts and which holds
for C-C bonds and the bond order n greater than one, can be applied to metal-metal bonds
and n less than one. Hume-Rothery [20] collected some results which contradicted
Pauling' s statements on this point. However, even if equation 28 were of general applica-
bility (as some modern authors assume [21]) a very mildly critical reader would still find
many questionable steps in the procedure leading to the table of valencies and 8 values.
1.1.3 The Engel-Brewer theory of metals and alloys
This theory has a number of features that are similar to the ideas of Pauling:
directed valencies, an important role of hybridization of orbitals on atoms constituting the
metal, widely changing valencies and the omnipresent electron pairs.
Brewer illustrates his theory with the example of tungsten [26], The configuration
of tungsten in a free atom ground state is d4s 2. However, the two s-electrons form,
according to Brewer, a closed shell, which is non-binding and which in the solid state
causes repulsion of other tungsten atoms. However, the configuration dSs is only 33,5
kJ/mol (8 kcal/mol) above the ground state, this difference being called promotion energy
of the d 5 s configuration, and the d4sp configuration is 230kJ/mol (55 kcal/mol) above the
ground state. Upon forming the metal, the energy of the das 2 configuration is supposed to
be lowered by 569kJ/mol (136 kcal/mol), the dSs configuration by 890kJ/mol (211
kcal/mol) and dnsp configuration by 569kJ/mol (136 kcal/mol). We shall now examine the
procedure by which the numerical values are obtained.
Following Hume-Rothery, Engel [24] associated crystallographic structures with
numbers of valence electrons in certain orbitals, i.e. with certain electronic configurations.
Having in mind the elements: sodium (bcc), magnesium (hcp) and aluminum (fcc), with
one, two and three valence electrons respectively, he suggested that the transition elements
with the configuration dn-ls should have a bcc structure, with dn-2sp they should have the
hexagonal close-packed structure and with dn-3sp 2 the fcc-structure where n is number of
valence electrons. Of course, some small deviations in n (for example alloys) are tolerated.
Vice versa, knowing the crystallographic structure one can determine the number and
distribution of the valence electrons over the orbitals. The authors of the theory [24-27]
assumed further that the contribution per s or p electron is given by the interpolation line,
which connects the points for metals having no binding by d-electrons, and serves as a
calibration (see figure 8).
The contribution to the binding strength by d-electrons is calculated in the
following way. The promotion energy is subtracted from the sublimation energy: the
former is fixed by the crystallographic structure of the metal in question. The structure
determines, namely, how many electrons should be in the s and p orbitals.
24 chapter 1
60
f i gure 8
*~ 50
Brewer-Engel theory of metals -3
E
Bondi ng energy (kcal/mole electron) c 40
0
of the indicated electrons (4 s,p or -~
3d, resp.) as a f unct i on of the position ~ 30
in the peri odi c table. El ement s of the
f i rst long peri od are shown. -a 20
u
ue
E,p (the upper curve) is estimated by I O
inter~extrapolation.
E a cal cul at ed as descri bed in the text. o
Co Sc Ti V Cr Mn Ire Co Ni Cu Zn
I I I l i 1 1 I I
XX
X
3d
F o = -
/ 1 . l l 1 1 1 t [
0 I 2 3 4 5 4 3 2 I 0
No. of u npair e d e l e c t r o n s p e r a t o m
The total contribution by s, p bonding is then subtracted, values being taken from graphs
such as that in figure 8, and the rest of the binding energy is divided by the number of
unpaired d-electrons. For example, hcp cobalt is expected to have the configuration dTsp.
From the sum of all d-orbitals, two should be occupied by pairs of electrons and three by
unpaired electrons. The maximum possible number of unpaired electrons is considered as
the ground state configuration. As can be seen from figure 8, while the contribution to the
binding energy by s,p orbitals increases monotonically with atomic number, the contributi-
on by unpaired d-electrons decreases. By circular argument, the authors [24-27] rationalize
the crystallographic patterns in the periodic table of elements, using values such as those
shown in figure 8. Sometimes the assignment of the most stable configuration appears to
be easy, as with molybdenum and tungsten, but in other cases various configurations lead
to very similar energies and thus to uncertainties, such as is the case with yttrium and
zirconium.
The Engel-Brewer theory has also been applied to problems of the stability and
crystallographic structure of alloys, in particular to structures of some intermetallic
compounds. Such compounds are formed when a metal on the left-hand side of the
periodic table (i.e. a metal with almost empty d-orbitals) is combined with a metal on the
right-hand side, where elements have several d-orbitals with paired d-electrons. Brewer
stated [26] that " the use of empty orbitals of hafnium and tantalum by the non-bonding
(i.e. paired) electrons of osmium or platinum could optimize the use of available orbitals
and electrons, and approach the optimal binding achieved by tungsten" . Using the example
of Hflr 3 Brewer illustrated how difficult it is to make quantitative predictions of heats of
alloy (compound) formation, that is, to go beyond qualitative predictions. Nevertheless, the
number of cases of binary and ternary alloys where the predictions are satisfying is
Structure and properties of metals and alloys 25
respectable.
Although successful as a semi-empirical approach, the theory [24-27] gave rise to
some serious criticism. The obvious problem [28] is how to believe any correlation based
on three outer electrons in fcc structures, when this is so far from the final description we
need for noble metals? The Fermi surfaces of copper, silver and gold clearly show the
presence of one sp electron per atom. Further, some low temperature structures are
probably different [28] from those suggested [24-27]. The most important problem is the
explanation of the structure of some magnetic elements and, on the other hand, the
absence of magnetism in configurations like that of copper [29]. Some conclusions
concerning alloys and structures stable at high pressures have also been criticized [29].
Modem experimental techniques (see chapter 3) have also made the assumption concer-
ning extended charge transfer in alloys such as HfPtx doubtful.
The Engel-Brewer theory has however been appreciated by some chemists [30].
The basis of the application was the idea that, by varying the composition of some alloys,
one can go from one crystallographic structure to another. For example, one can start with
pure molybdenum of the bcc structure, dissolve increasing amounts of iridium in it, until
at a concentration known from the phase diagram, the structure switches over into the fcc
structure of iridium. It means that in the state before alloying the elements had to have
one sp-electron on the Mo-rich side, and three valence sp-electrons on the fcc side of the
phase diagram. However, the change in the number of the sp-electrons leads according to
the Engel-Brewer theory to a change in the number of d-electrons. Thus, it was expected
[30], that the number of d-electrons could be varied by changing the composition of the
alloys. It was not appreciated in this approach that the Engel-Brewer theory makes an
assumption about the electronic configuration of the free atoms from which the alloy is
made. The Engel-Brewer theory does not draw any conclusion about the electronic
structure of the solid alloys and of course says nothing about the surface composition of
alloys.
1.1.4 The Miedema theory of stability of alloys
This theory starts with a cellular model of solids [31]. A crystal of an alloy is
divided into cells by planes which bisect the distances between the nearest neighbours and
form the so-called Wigner-Seitz cells, which are analogous to the Brillouin zones in the
reciprocal space, as discussed in 1.1.1. In an alloy, cells around elements of different
electronegativity have different densities of electrons on their boundaries. Miedema
suggested [32] that the enthalpy of formation of an alloy can be calculated by an empirical
equation:
AH = f(c) [- P.e (Ate*) 2 + Q (Anws) 2] (31)
26 chapter 1
where P and Q are constants, f(c) is a symmetrical function of the molar ratios, and for an
alloy AB forming a solid solution it is XA(1-XA), XA being the mole fraction of component
A. A~)* is the difference in the values of ~* for the two elements, ~* being to a first
approximation the work function ~; Anws is the difference in the values of electron
densities in the Wigner-Seitz cells, all corresponding to A and B, respectively. Miedema
showed that a more self-consistent system of enthalpies of formation, in better agreement
with values known from experiment, can be obtained if one uses the adjusted ~* values
tabulated by the author. The difference between ~ and ~* is small for platinum (5.55 vs
5.65 V), but somewhat large for some other elements (for Zr, 3.15 vs 4.05 V). Miedema
suggested calculating the densities nws by using (B/Vm) v~, where B is the bulk modulus of
compressibility and V m the molar volume.
The idea behind equation 31 is that electrons are transferred from atoms of a metal
of lower electronegativity to atoms of a metal of higher electronegativity. According to
Miedema [32], the charge transferred per a t om mT~ a can be calculated by
AZ A = 1.2 (1-XA) A~* (32)
This means that in Hflr 3 about 0.7 of an electron per hafnium atom is transferred from
hafnium to iridium. The Engel-Brewer theory, which also explains the high stability of this
compound (see 1.1.3) , assumes an opposite electron transfer. The experimental results,
e.g. core level shifts, on various compounds of this type indicate that most likely there is
no electron transfer at all (see chapter 3), but formation of strong partially-localized bonds
takes place between unlike elements (see chapter 2).
The practical success of this theory is indisputable. It is almost impossible to check
the stability experimentally and to make some predictions concerning phase diagrams of
all alloys of potential interest for material sciences. Miedema' s theory, however, offers a
certain tool for making rough but useful predictions, where experimental results are
lacking.
The theoretical background of the theory is however weak. It is too strongly
associated with the assumed charge transfer between the components of alloys, and moreo-
ver, while the work function ~ is indeed a measure of the electronegativity of metal
surfaces, a substantial contribution to ~ is made by the surface dipole, which is not present
in the bulk at the Wigner-Seitz cell boundaries, where the charge transfer should take
place.
1.1.5
The quantum theory of alloys
Quantum mechanical calculations on small organic molecules can achieve a very
high accuracy, which is impossible to achieve with large systems of interacting particles,
Structure and properties of metals and alloys 27
such as solid crystals. Yet the fact that the potential in the solid can be taken as periodic,
and Born-Karman conditions can be assumed to be fullfilled (see 1.1.1), allows us to be
somewhat precise when treating the properties of large single crystals of metallic elements
(see for example a comparison of calculated band structures with those derived from
electron photoemission in chapter 3). However, when a random alloy is formed with
elements A and B, the mole fractions being x A and xB, the periodicity of the potential is
abolished and the degree of sophistication which is needed for a description of the same
accuracy is considerably enhanced.
Faulkner summarized the early development of the quantum theory of alloys in a
paper [33] which we shall follow.
The potential in the alloy A-B at a point r can be written as a sum of contributions
from different lattice sites (Rn):
V(r) : n Wn ( r - Rn) (33)
V n is V A or VB according to the atom on site n, but the fact that in random alloys V is no
longer periodic is a problem in the description of alloys. There are several ways of coping
with this difficulty, but we shall mention only three of them.
(1) I n the Rigid Band Theory (RBT) one neglects the difference between A and B and
assumes that the only consequence of substituting A for B is that the common band is
occupied to a higher or lower degree, viz. E F is shifted, just by adding electrons to or
extracting them from the pool of electrons under the Fermi surface [34,35]. A consequence
of this model is that charge is freely transferred from one component to another, for
example, from copper to nickel. The RBT was for a long time the basis of early theories
of catalysis by alloys [9,19,36], but the total failure of this theory, and of ideas behind it,
to explain the photoemission results (see chapters 2 and 3) stopped its application after
about 1968, when the papers by Spicer appeared [37].
(2) The next level of approximation is a model of a virtual crystal with an average
potential VAV [38,39] on each lattice point:
WAy = XAVA(r) + XBVB(r) (34)
The difference V Av(r ) - Vo(r), where Vo(r ) is the ideal periodic potential, can be treated as
a perturbation and it leads to small deviations from the Eo(k ) function for the periodic
potential. It has been shown that this is also a rather poor approximation.
(3) Higher approximations stem from the theory of multiple scattering phenomena. This is
appropriate, because the crystal orbital ~ is, in the context of these theories, constructed in
such a way that the delocalization of electrons outside the atomic spheres is formally
described by wave functions which look like a combination of " incoming" waves with
28 chapter 1
waves " scattered" by surrounding atoms. Atoms A and B are in this way considered as
unlike scatterers converting the incoming function into different scattered functions, by the
operation of potentials V A and VB. The operator which relates the incoming and scattered
waves is t, there being different values tA and tB for each type of atoms. In early attempts
an averaged scattering operator was used.
t av = XA tA + x B tB (35)
but the results were even worse than with the approximation of the virtual crystal. The
break through came when Soven [40,41] suggested using the following picture. A virtual
crystal is constructed which has an initially undetermined coherent potential W(r) on each
site. The scattering is caused by local deviations from this potential, so that the scattering
operators are:
tA = (VA- W) + (VA- W) CJ tA
tB = ( VB - W ) + ( V 8 - W ) CJ tB ( 3 6 )
In equation 36,
further below.
stands for the so-called Green operator, which will be briefly discussed
The reader is already familiar with the Schr6dinger equation"
[-h2/2m) V 2 + V (r)] ~ = E~ (37)
which in the operator form reads as"
12I~ = E~ or ( E- 121) ~ = 0 (38)
The Green operator is defined by an analogous operator equation:
( E- 121) G = 1 (39)
and it is very useful in describing scattering phenomena or other quantum mechanical
problems, such as the construction of wave functions for crystals, which have a similar
structure. For example, in the formalism and the language of the multiple scattering, the
solution of equation 37 is written as [33]:
V- - r -I" (~Jo ]~n tn ~]/~n (40)
Structure and properties of metals and alloys 29
where ~n is constructed in the form of
~n = (~ + Go ]~m,n tm 1]/im (41)
The function ~) represents the incoming wave on the system and /I/in is the total incoming
wave on the site n, that is, including the scattered waves coming from other sites m, and
G O the Green operator of the free propagation of the wave ~.
The procedure of the Coherent Potential Approximation (CPA) is to determine the
potential W from the condition that
XAt a + X B t B = 0 (42)
This condition means, in the language of scattering, that the electron propagates in the
alloy as in the virtual crystal and experiences scattering by V A and VB (equation 36), but
scattering effects cancel on average (equation 42). After W is determined, the properties of
the alloy can be found by manipulations of Green operators and its matrix elements. The
most important of these manipulations is the calculation of the density of states function
from the equation
N(E) =- l m. Z m Im < Gmm> (43)
In this equation, Gmmis the matrix element of the operator G, with orthogonalized
functions localized on site m. The term Im indicates that only the imaginary parts of the
matrix elements are used in equation 43. By studying various restricted averages of the
matrix elements of the Green function, one can determine various local densities of states,
for example, the average state density localized around a particular type of atom [41]. The
results of such calculations for some interesting systems are shown in figures 9 and 10
[42,33]. By comparing the results of calculations with the experimental photoemission
results [33,42] in figs.9 and 10, we can see that the CPA theory successfully describes
their main features. The basis for success lies in the theory properly acknowledging that
sites A and B can differ substantially not only in number of valence electrons but also in
respect of V A and VB. When inspecting figs. 9 and 10 one can easily arrive at the
conclusion that, when a molecule like carbon monoxide interacts with a specific site on
the surface, it feels whether that site is a nickel or a copper atom. According to the Rigid
Band Theory or the virtual crystal theory, it would not be so, because in those models the
individual atoms would be indistinguishable.
30 chapter 1
5 --1
7 7 % Cu, 2 3 % N, |
o ? . _ ,/ h x \ -
z f ,/ I -\ \ \ - , ~, _
i / / , ' " \ _
o /
i . . . . . . r - . . . . . " l
- 0 . 7 - 0 . 6 - 0 . 5 - 0 . 4 - 0 . 3 - 0 . 2 - 0 . I 0
ENERGY BELOW E f (r yd be r g s )
4O
35
>~
3O E
o
25 ~
o
2 0 ~
i---
~5 ~
LL
I 0 o
> -
5 ~
Z
0
figure 9
Density of states f or a Cu-Ni
alloy and the projected densities
of states f or the Cu and Ni sites.
Notice: the UPS experimental
results are at lowest energies
deformed by the artefacts of the
experimental techniques but the
region round E I is correctly pro-
bed [44b1.
figure 10
Comparison of XPS valence-band
spectrum f or an evaporated Cu-Ni
alloy sample with a density of
states f or Cuo.6-Nio.4 calculated
in CPA. (from ref 42)
I 0 ! ! Q i I .~
/
/
.1
0.5 #
o
0 1 9
8 6 4 2 0 - 2
BI ND I NG ENER GY ( e V )
The Coherent Potential Approximation (CPA) has been worked out in great detail
[43,44] and further modified and extended to ordered alloys. It goes much beyond the
scope of this book to discuss these developments, but we refer the interested reader to
some selected papers [45]. An extended comparison of experimental and theoretical UPS
and XPS intensities has been presented [46].
Structure and properties of metals and alloys 31
Theoretical calculations using the CPA [47] and experimental XPS results [48]
have been compared and found to agree for the Pd-Hx system (foreign atom is placed
interstitially); the experimentally found hydrogen-induced states centered at 5.4 eV below
the Fermi level strongly support the idea that hydrogen is present as atoms and not as
protons which have injected their electrons into the d-holes of palladium. Interstitial alloys
behave probably in a similar way.
In relation to the various semi-empirical theories of alloys (see sections 1.1.3 and
1.1.4) we note that the alloys of s-metals have been also analyzed by CPA theory [49].
For these alloys the CPA theory predicts the existence of various localized states in the
whole range of energies of valence electrons.
There are also some other sophisticated methods which either generally or for some
special cases are still better suited for exact calculations than the original form of the CPA
theory [50,51 ].
Alloying is known to cause some redistribution of electrons on the individual
atoms of alloy components. For example, palladium in silver has a narrower d-band than
in pure palladium. In consequence, from a certain dilution up (about 60% silver), the
whole narrowed band of states localized around palladium atoms falls below the Fermi
level. This band therefore becomes fully occupied at the expense of the s-band. The
narrowing of the d-band results from the diminished overlap of the palladium orbitals (see
section 1.1.1 for the relation between overlap and band width), and from the suppressing
of the d-d electron repulsion by dilution. These and similar effects should also exist in
some other alloys and one has always to consider the possibility of a change in electron
configuration caused by alloying. More intriguing is the question to what extent charge
transfer between alloy components occurs. In chapters 2 and 3 we will discuss some
results dealing with this point, but first we consider attempts to deal theoretically with this
problem.
Kfibler et al. [52] used the Augmented Spherical Wave method and the local
functional density theory for self-consistent calculations; they calculated the total and
partial density of states, that is, expressed per component and per orbital of a given
symmetry, and from that they determined the occupation of s, p and d orbitals of
individual components. They concluded, for example, that in Zr3Pd the palladium atoms
receive 0.6 s,p electrons per atom from the zirconium atoms which each lose 0.2 s,p elec-
trons per atom. The intra-atomic transfer of s to d electrons on palladium is calculated to
be nearly zero. A closer inspection of calculations made by the same technique [52]
reveals that the configurations for pure metals do not agree with the experimental results
(e.g. copper has 9.5 d-electrons, instead of 10), so that the predicted charge transfers in
alloys might be doubted. Predictions for core level shifts are also presented [52].
Early studies of the M6ssbauer effect on alloys of transition metals revealed
somewhat large isomer shifts which were at first explained solely by charge transfer
32 chapter 1
between the alloy components (see chapter 3). However, later thorough theoretical work
revealed that the original straightforward explanation needed serious corrections [53].
When an atom A with a spatially extended p- or d-orbital is squeezed in a lattice of
another element B without that particular feature, electrons on the far reaching p-orbitals
simulate in the space around B atoms a charge transfer. An analysis shows that indeed the
p-like charge increases on B, but that this is mainly due to the tailing of the p-orbitals
from A into the B atomic spheres and not to an increase of p-electrons on the on site
orbitals of B, or to other bonding effects. Attempts were made to subtract the tailing effect
or pseudo-charge transfer from the total charge transfer, the latter being calculated by
integrating the density of electrons within the Wigner-Seitz or atomic spheres. Results of
these very delicate calculations are shown in figure 11.
o l ' ' ' ' ' ' ' I . . . . ' ' t < 1
Tail Onl y _ / Mod Mul hken i
o.o , . . . . . . . . . . . . I . . . . . . . . . . . . .
_
A - i u Compound s o
Pt Compound s []
Ir Compound s , ,
It . Pt , and Au
0. 2 Sit es ri, . ~. . ,
o, \ \ _
O0 - j
-0,1 +
Hf Ta W Re Os Ir Pt HI Ta W Re Os Ir Pt
-01
-02
figure 11
Residual charge transfer at Ir,
Pt and Au and the other atomic
sites in compounds of the CsCI
structure compounds after the
effects of tailing were subtrac-
ted. This involves taking the
total charge transfer and sub-
stracting of the tailing charge.
The right hand column shows
the results f or the residual char-
ge when the modified Mulliken
scheme is used to estimate the
tailing. The left hand column
shows the results when overlap
contributions to the tailing are
neglected and is shown to provi-
de some sense of how sensitive
the results are to the treatment of the overlap. The open symbols were obtained using
elemental volumes, however, when hafnium compounds have volumes a f ew percent
smaller than the sums of the elemental volumes and the solid symbols indicate the
consequences i f this volume contraction is assigned to the hafnium site. ~rom ref 53)
Structure and properties of metals and alloys
33
For catalytically interesting alloys, for example platinum-gold, the charge transfer
to platinum is +0.2 electrons when correction for tailing is not made, but with this
correction it is -0.1. The difference in calculated charge transfer values, which is only the
result of the use of different methods of calculations can be as large as 0.5 electrons.
Conclusions are therefore very cautiously drawn [53]; too much should not be read into
our figure 11 and perhaps of greatest interest is that once charge tailing is accounted for,
the remaining changes in electron counts are consistent with a picture where electronegati-
vities increase as one traverses the 5d row from hafnium to the elements to its right, with
gold however being less electronegative than platinum and perhaps iridium as well. While
consistent with some notions this is inconsistent with many notions concerning the
chemistry of gold" . The difficulty in assessing theoretically the extent of the charge
transfer is clearly and simply demonstrated by this: one has to calculate charge transfer of
the order of 0.1 electrons per atom with up to 10 electrons being involved, not knowing
exactly over what space one has to count the electrons belonging to each component.
In chapter 3, several electron density contour maps are presented (figures 13,21,22)
which also touch the problem of the charge transfer. Actually they do not show much that
one would call charge transfer. The reader will also find in chapter 3 other views and
preliminary conclusions on the problem of charge transfer as derived from the totality of
all results presented by that chapter.
Finally, we mention the electronic theory of ordering and segregation in catalytical-
ly less interesting alloys such as these formed between simple metals, for example, alkali
and noble metals [54]. An important step is made by using microscopic quantum mechani-
cal calculations to predict macroscopic thermodynamic behaviour. Values for charge
transfer between alloy components are derived, for example, for the 1:1 alloys: Li-Cs, 0.25
electron/atom from Cs to Li; Na-Cr, 0.27 electron/at from Cs to Na; Cu-Au, 0.11 electron
per/ at from Cu to Au; Ag-Au, 0.13 electron/at from Ag to Au.
We have seen above how difficult it is to predict, if only in a qualitative way, the
main features of alloys; use of the coherent potential approximation (CPA) theory was the
first real breakhtrough. To predict the properties quantitatively, e.g. the amount of charge
transfer, is again a task an order of magnitude more difficult. However, catalytic chemists
want to know the composition and the electronic structure of alloy surfaces as well as
their properties in chemisorption and catalysis. It is an extremely demanding task, but
some progress has already been made in this direction [55-57]. For example a prediction
has been made concerning the binding energy on a surface of an nickel-copper alloy [56].
Atomic adsorption on the hollow square of four atoms is considered and the conclusion is
reached that the binding energy on this cluster is so much influenced by the average
composition of the alloy that it drops by a factor of two when going from pure nickel to
very dilute nickel-copper alloys. However an effect of this size seems to be at variance
with the experimental results mentioned in other parts of this book.
34 chapter 1
The problem of adsorption on alloys has been also approached theoretically by
other authors, for example, in the book by van Santen [59], whose treatment is simpler
than that in [56] although it goes further in applying the theory. Van Santen concludes that
there is an electronic structure effect on the binding energy of hydrogen atoms on an alloy
cluster where the average number of valence electrons in the alloy is a variable.
The nickel-copper system is a favourite subject for calculations, since it concerns
somewhat light elements, and also many of the experimental results relate to it. Castellani
[60] studied it by Extended Htickel Theory and concluded that there should be a charge
transfer and an effect of copper on adsorption properties of nickel. Placing of nickel into a
copper matrix causes according [60] a decrease of about 30 kJ/mol in the heat of
adsorption of carbon monoxide, when compared with nickel in a nickel matrix. An even
more pronounced charge transfer from copper to nickel was found in [61 ].
Simple LCAO theories are good enough to explore new phenomena [59-63] such
as adsorption on multicomponent systems, and to introduce the terms necessary for the
description of the experimental results, but they are not very reliable for making quantitati-
ve predictions. However, some pioneering work in higher approximation has already been
carried out too. Muscat [64] has studied hydrogen adsorption on the (111) face of copper,
in the surface of which a nickel atom impurity is placed. Binding energy was first
calculated for a cluster using 19 muffin-tin potentials of either pure copper or with one
nickel atom instead of one copper atom (see figure 12).
f i gure 12
Model clusters HCu19 or HCu18Ni used in
calculations by Muscat. Ci rcl es-upmost layer,
triangles - the layer under it, square - an at om
in the next lower layer. Cross indicates the
position of H. Placing Ni in posi t i on 1 changes
the one electron energy of the system by 0. 40
a.u. (in other words the ensemble NiCu2 of
nearest atoms behaves very differently f r om the
ensemble Cu3). However pl aci ng of Ni in posi-
tion 2, 3 or 4 causes an effect 4-5 times smal-
ler and with Ni in 5 and 6 the effect is zero
(this is a negligible interaction through the
metal).
In the following step, the energy change was calculated due to the cluster being embedded
into an effective medium consisting of a homogeneous electron gas. The results were quite
interesting: the effect of introducing a nickel atom in a copper matrix is only important
when a nickel atom is in the position 1 (see figure 12). In position 2 it is five times
Structure and properties of metals and alloys 35
smaller and in position 3 it is small and of the opposite sign; in position 5 it has no
influence at all. Obviously, hardly any of the effect of nickel is through-the-lattice, viz.
through the collective metal properties; it is clearly a short range, chemical bond effect.
We can extrapolate this conclusion and say that on the (111) surface of copper-nickel
alloys one can expect four types of tri-atomic cluster each showing a distinctly different
binding energy towards hydrogen: Ni 3, Ni2Cu, NiCu2, Cu3 , the last showing very weak
binding; we may then try to explain chemisorption and catalytic results by this model, in
which the ensemble size (Ni 3, Ni2 ...) plays the most important role. We shall turn to this
point in chapter 9 and shall see that this approach can be successful.
The results for the calculated charge transfer and ligand effects show an interesting
and clear picture: the simpler the theory is, the more pronounced is the charge transfer.
The reader will find further discussion on this point in chapter 3 and elsewhere in this
book.
1.2 Some results of the theory of chemisorption on metals and alloys
1.2.1 General features of chemisorption - a qualitative picture based on quantum chemical
calculations
1.2.1.1 Chemisorption of atoms
The reader will be familiar with the quantum mechanical theory of bonding in a
molecule A-B. The simplest case is that of two atoms each having one valence electron in
one atomic orbital (see [1-4] or, more advanced text [65]). The main terms and results of
such a theory are summarized in figure 13 and the information contains the first ingredient
needed to build up a qualitative picture of a chemisorption theory. A solid is actually a
giant molecule. Due to the mutual interaction of all atoms in the solid, the molecular
energy levels form a whole band of levels (see 1.1.1 and figure 14 below). In analogy
with diatomic molecules, the lower part of the band is called bonding and the upper one
antibonding. We have presented a simple theory of band formation in section 1.1.1 from
which we know that the band width is proportional to the matrix element 13 (1.1.1,
equation 19).
In principle, both the metal atoms, on left-hand side of figure 15, and the adsorbed
atoms or molecules at high surface coverages can form a band of energy levels [66].
However, let us start with the simplest case. A single atom, with one electron in a single
valence orbital on a single energy level, interacts with a solid, the electrons of which
occupy a band of energies (fig.15, left-hand side) The interaction of an atom with a solid
can be described by one of the two limiting cases:
(i) the interaction is weak and leads to a broadening of level A into a virtual band
3 6 chapter 1
( ii)
i nsi de the metal band;
the interaction is strong and can be descri bed as the formati on of a ps eudomol ecul e
from atom A and one or several surface atoms of the surface; by interaction with
the solid, the energ y l evel s of the pseudomol ecul es are broadened ag ain into a
narrow band.
Be f o r e bond f o r ma t i o n Be f o r e bond f or ma t i on
. E
b b ~
me t a l a t om
~B "
a mpl i t u d e s
~a b "
/
/
/
/
/
\
\
a b
I EA B
/ ~ a ba nt i bond i ng M. O. \ \
/ I \
/
/ I /
/
I /
AE /
/
I /
, z
/
I /
\ \ I Eb B
~B' b o n d i n g M.O.
a f t e r bond f o r ma t i o n'
b a
b a
Ea, ~ a
a d s o r ba t e at om
f i gure 13
Lef t and right: energy levels of atoms b and a, before a molecule AB has been f ormed.
In the middle, energy levels (bonding-B, antibonding ab) corresponding to the mol ecul e
AB are f or med by the interactions of electrons on the level b and a. In the crudest
approxi mat i on:
2 89
A E = [ 4 V 2 , b , a + ( E b - E a ) } , with Vo, a = ~dO o n ~ dr,
E a = ] , / 2 ( E b + E a ) + { V 2 b , a + 1/4 ( E b - E o ) 2 } 89
Structure and properties of metals and alloys 37
a) b ) c) - - - - ~ N ( E)
o ~
N=8 N= oo N= oo
figure 14
One dimensional chain of atoms, with indicated numbers of atoms N
a) b) - energy bands
c) - density of states curve, corresponding to b)
l
The theory for both limiting cases was developed in the 60' s, and in some earlier
pioneering papers [67]. What follows is based on some of the original papers [68-71] and
some reviews [15,16,59]. Let us start the discussion with the weak-bond limit.
figure 15
Weak bond limit in the formation of a bond between atom A (single orbital, single
electron) and atoms of metals. Metal electrons occupy the energy band up to Fermi
Energy. By interaction, level A broadens and becomes partly occupied.
The interaction of an electron in a n E A level (figure 15, right) with electrons in the
band (figure 15, left) leads to two effects on the E n level:
(i) the E A level is shifted on the energy scale, and
38 chapter 1
(ii) by the interaction with atoms, the electrons of which form the band, it is broade-
ned.
The broadened level E A can be fully or partially occupied by electrons or be completely
empty, in which case it represents an A + ion. When the energy band for the metal is
narrow, the broadening of the adsorption level is less pronounced than when the band is
broad.
Let us now make a step to the strong-bond limit and consider the so-called
pseudomolecules. The strong pseudomolecular interaction splits the E A level into two
broadened levels: a bonding and an antibonding one, with a gap in between. This is
similar to the situation with the molecule AB in figure 13. The antibonding band can be
either above or below the Fermi level E F, or it can be split by EF into an occupied and
unoccupied part. This last case is shown in figure 16.
d 1
S
be f or e
i nt e r a c t i o n
- - EF~
E A
E
- E A -
N(E)
me t a l ,
o nt i b o nd i ng
EA- me t a l ,
b o nd i ng
f i gur e 16
St rong chemi sorpt i on ( pseudomol ecul es) limit. Chemi sorpt i on of at om A on a t ransi t i on
met al with s and d bands. Lef t - bef ore interaction, right - af t er interaction.
We shall now investigate whether and how the simple theory can explain the trends
in the chemisorption bond strength for adsorbed atoms when metals of the periodic table
are compared.
Theoretical analysis [15,16, 59, 68-71] shows that we can make the following
simple statements.
(i) The position of the E A levels (bands) and their separation from each other are mainly
due to the interaction of the electron in the EA levels with those in the orbitals of the
solid. It is mainly a group of d-orbitals on the metal atoms adjacent the A atom which
have, moreover, suitable symmetry that are involved in forming a bond with A [72,73].
These d-orbitals are sometimes called ' group' orbitals.
Structure and properties of metals and alloys 39
(ii) The interaction of the E A electrons with those of the sp-band is mainly responsible for
the level-broadening, while its influence on the downward movement of EA is less
important. The latter aspect should be particularly kept in mind when the metals of one
period are compared with each other: the variations in the number of the sp-electrons ns
are smaller than variations in the n d, the number of d-electrons.
The exact position of the broadened levels EA also depends on the strength of the
electron-electron interaction in the given orbitals. These considerations are sometimes
thought to be sufficient to explain the trends in chemisorption bond strength along the
periods [15,16,59,68-71]. When going from left to right in the Periodic Table, the band
width and its occupation changes as is roughly illustrated in figure 17.
T i Fe Cu
I I
I I
I
F
" ~Z
f
figure 17
Position and width of the d-band as a function of the atomic number Z, within the first
metal period of the Periodic Table (the occupied part is indicated).
The trend in figure 17 is caused by the effect of nuclear charge (Z e): a higher
nuclear charge contracts the orbitals and lowers the overlap, so that a lower band width
results; the band narrowing and decrease in the energy of the band occur in the conse-
quence. Certainly, the greater the nuclear charge, the lower is the energy of the electrons
in the bands. A similar trend is observed in the positions of both the antibonding and
bonding E A levels (bands). The antibonding band is almost empty for metals on the left of
the Periodic Table, but becomes increasingly occupied as Z increases. The more electrons
that occupy the antibonding bands, the higher is the total energy and the lower is the
chemisorption bond strength. This explains why the heats of dissociative adsorptions of
hydrogen and oxygen, and so the respective atomic chemisorption bond strengths, decrease
40 chapter 1
on passing from left to right across each period [68,69]. The trends are shown in figure
18.
f i gure 18
Experi ment al l y measured
chemi sorpt i on energies >
f or hydrogen and oxygen "-"
on the 3d transition metals. >-
(_9
CV_
(selection by [68,691). w
Z
The trends are the same in tu
the 4d and 5d series.
z
The theoretical results o
shown are f r om a model a_
t2s
calculation within the o
t.t3
effective medi um theory ~_
I11
[68,691. I
( D
~176 1
- 2. 0
- 4 . 0
0. 0
- 4 . 0
-8.0-
- 1 2 . 0 -
9 THEORY
[ H Y D R O G E N l
a E X P E R I M E N T
9 ~ ~ 9
I I I 1 I I ! 1 I
[ OXYGEN]
r l
9 T HEORY
u EX PER I M ENT
( p o l y )
! " 1
r l
~ T - - ' i I i I I i I i
Sc Ti V Cr M n Fe Co Ni Cu
The theory predicts that, in all cases of adsorption of atoms, the most likely
position of an adsorbed atom is in the valley between several surface metal atoms. The
higher the coordination of the adsorbed atom, the greater the energy lowering by adsorpti-
on. Adsorbed atoms prefer to occupy the sites at which the crystal would grow by
accretion of further metal atoms. With non-dissociating molecules the situation is different,
as we shall see below.
Several authors have addressed the problem of the localization of the chemisorption
bond. Does an adsorbed atom on one site influence the adsorption of another atom on an
adjacent site by an interaction through the metal? This problem also exists with regard to
the promoters or supports interacting strongly with metal particle. The answer is in the
affirmative; there is indeed an interaction through-the-metal [74], but the size of this
interaction requires discussion. Schrieffer approached this problem by analysis of experi-
mental results concerning the strength of the interatomic interaction of chemisorbed atoms.
He concluded [75] that the total of through-the-vacuum and through-the-metal interaction
Structure and properties of metals and alloys 41
energy amounted to about 5% of the energy of the chemisorption bond formation. It is not
known how much of this is due to the through-the-metal interaction, but in any case 5% of
the adsorption energy would be the limit. This all indicates that any through-the-metal
interaction is of a very short range character. This conclusion is also supported by figure
19 [76], where we can see that the change in the electron density due to an adsorbed
hydrogen atom is very limited in space.
1
I I I I ; / ; ; / I I I I I ~ 8
4
1 1 i i ! 1 1 0
-8 -4 0 4 8
f i gure 19
Total electron density map, H-at om on 7el l i um' -met al (paramet er r s = 2) Haa s is in its
equilibrium posi t i on of 1.1 a.u. f r om the j el l i um edge. The shaded area indicates the
posi t i ve background of 7ellium' . Distances in atomic units (a.u. = 0.053 nm). For
compari son the lattice constant of Ni is 6.66 a.u.; the paramet er r s = 2 describes the
somewhat high density of the positive charge, f or example in a metal such as Al or Pb.
1.2.1.2 Chemi sorpt i on of undissociated molecules
The main features of the theory dealing with this problem can be best understood
by taking adsorption of carbon monoxide as an example. Metals can be classified into the
following groups.
(i) The block of metals between Sc to La (on the left side) and Cr to W (on the right
side). Adsorption of carbon monoxide is dissociative at low temperature and the
resulting oxides are stable against reduction by hydrogen or carbon monoxide; Mn
and Re probably behave similarly.
42 chapter 1
(ii) Fe, Co, Ni, Ru, Rh. These metals dissociate CO at room temperature such as Fe or
slightly above 400 K such as Ni. Oxides are reducible by CO above 500 K.
(iii) Pd, Ir, Pt. Carbon monoxide is adsorbed strongly, but non-dissociatively, unless the
surface is highly defective [77].
(iv) Cu, Ag, Au, Zn, Cd and Hg adsorb CO very weakly.
We shall discuss below the non-dissociative adsorption of CO, but many features of this
adsorption are also common to other molecules: NO, ethene, alkenes in general, benzene,
etc.
In the usual approximation, orbitals of free atoms which constitute the molecule
combine to form molecular orbitals. A correlation diagram showing which atomic orbitals
form which molecular orbitals in the case of CO is presented in figure 20. Even the
simplest theories give us a reasonably accurate idea about the form of the electron
densities, viz. the shape of particular molecular orbitals. A very useful collection of shapes
of molecular orbitals has been presented by Jorgensen and Salem [78]. The shape of those
orbitals which are most relevant for chemisorption of CO (Frontier orbitals, including the
highest occupied and lowest unoccupied orbitals, i.e. HOMO, LUMO) are shown in figure
21 [78].
Electrons of the metal interact through the orbitals of the surface atoms (in
particular the group orbitals, see above) with the bonding and the lowest-lying antibonding
molecular orbitals of the molecules. The situation for CO is as in figure 22. Such picture
was first used by Dewar to explain bonding in n-complexes of ethene, and was later
adopted by Blyholder to describe the chemisorption of CO [79]. The bonding due to shifts
of electrons indicated by the arrows occurs through what is called a donat i on- backdonat i -
on mechanism. While delocalization of the 5cy-electrons in the HOMO stabilizes the C-O
bond and frequency of the stretching vibration increases, partial occupation of the antibon-
ding orbitals by electrons destabilize it. The delocalization of the 5or-electrons is often
away from the M-C region as found e.g. for palladium and copper [80].
Occupation of the antibonding rt-orbitals of adsorbed CO is increased when an
positive ion is placed near to oxygen atom or the centre of the C-O bond. This is an
electrostatic and not a binding effect, since placing a point charge without any orbital has
the same effect [81].
When temperature is increased, more electrons can be raised into the rt*-antibon-
ding orbitals. A higher temperature also activates the tilt of the C-O axis [82] away from
the most stable orientation [83] and towards the metal surface. The perpendicular
orientation the most stable on (100) or (111) fcc faces, can also be perturbed by CO-CO
interactions, as is clearly the case on (110)fcc planes.
Structure and properties of metals and alloys 43
;~ff--W_,"2:;;
" h 2 , . ' . . . . .4
2/7, E: O. 1 2 6 8 "k. L. - ' - - - ; - ~ "
,',, - . . . . . . . . . . . . -'r . . . . ; '
. s ; , . ~ -, - . . - i - -
.~:.~-' -.- . ," 5 0 E=- O. 5 5 4 4
1 , E=- 0 . 6 3 9 5 ,, . ; } ; ' _- i . / . . . o
/ , . . , , . _ . . , _ . . _ . - , x, , . . - , . 1 IT E=- O, 6 3 9 5
V. ; I - - 7 - - , ' . . " . - "
40 E=- O. 8 0 3 8
3 0 F=- 1 . 5 2 1 0
figure 20
Shapes of selected MO's of CO [78].
44 chapter 1
6o- "
\ \
/ \ \
i 1 ~ / \ \
I I \ \
\ \
I I \ \
I I \ \
I I \ \
I I \ \
I / \ \ \ \ \ \ \
/ //
/ 27r ~
/ / . ~ , , . \
2 p - - e % - ~L _ _ a \ \ \ \
, ~ , " >- - ' 4 ~ - ~ 1 = . x 4 ~ - = . - >~ " ' * ' *
/ . - ~ . . . . ~ " + ' r < \ " \ ~"
\ \ , " -~,/\ \ \
\ \ . \ \ I \ \ \ \
\ \ 2 S
\ \ \ i
,,~ ~, 1
3 n
C CO 0
1, ~ ~0~~0 _. . . . . - - Q~~O so.
2% _ ~ . . . . . . . . . ~ ,0.
, % ~ . . . . . . . . . . @ - ~o
1 0" - - - - - - - - - - -
g " " --- -.. _ . . .
" " " - : Q 10"
N2 CO
figure 21
Correlation of orbital diagrams
Upper part: relation between atomic (C,O) and molecular (CO) orbitals, only orbitals of
higher energies shown
Lower part: a comparison of homo-atomic (N2) with hetero-atomic (CO) molecule.
N.B. The vertical position is related to the total energy, but is not on scale/
Structure and properties of metals and alloys 45
figure 22
A schematic picture
indicating the main
contributions to the
M-CO bonding.
I k
\\ V ~ ii
When the oxygen atom comes near to the surface, the molecule can dissociate. Whether or
not the dissociation takes place depends on the activation energy of the tilt and the
thermodynamics of the C and O adsorption [84-86]. These atoms have to be bound to
several contiguous surface atoms, that is to say to an ensemble, as has been established
both experimentally [87] and theoretically [88]. The theoretical calculation revealed that
various pathways of dissociation are often possible, but that they differ in activation
energy. Several of these pathways are shown in figure 23 [88].
v( ' ( " ( 111 ) 44 kcal moi - l
v( ' ( (10()) 36 kcal mol -~
/
figure 23
Minimum dissociation energy paths of
CO on large clusters of Rh. The activa-
tion energy on the (111) surface is
predicted to be 8 kcal moU (33
kJ mol 1) higher than on the (100) sur-
face.
Adsorption of alkenes is very similar to that of carbon monoxide, but the donation
mechanism is more important. With other multiple bonds some caution is necessary. For
example, one often sees a schematic picture for the bonding of ketones and aldehydes in
the following form:
46 chapter 1
\c o
J
" M
However, the position of the HUMO' s and LUMO' s of molecules having a C=O bond
[78] shows us that the donating HUMO orbital is localized on the atom O; in a classical
chemical terminology, it is the lone pair. It is then better to write schematically
C=O~M.
The geometry of this complex is not favourable for the back-donation.
Within the metals of groups 8-10, the heat of adsorption of carbon monoxide does
not vary very much, but, as with other molecules or atoms, it decreases when going from
the left to the right in each period [89]. There are clearly observable differences between
the various crystallographic planes of the same metal and we must remember this when
analyzing trends in chemisorption bond strengths. Calorimetric in measurements of
adsorption heats with molecular beams and single crystal planes [90] has shown that each
of the various surface structures, characterized by LEED and other means of surface
crystallography, has its characteristic adsorption enthalpy. Older calorimetric results,
obtained with polycrystalline materials such as powders and films, and quoted elsewhere
in this book, agree reasonably with an averaged value at least in this case. Thus, old
results on films and powders should be used with caution, but in the absence of a better
alternative they can be used for correlations with other physical properties of metals.
It is difficult to predict theoretically the optimal location for CO molecules to
adsorb. The differences between various potential sites are not large and the optimal
position is a result of various factors such as [59] (i) optimal binding with various group
orbitals, (ii) mi ni mum repulsion interactions between the fully occupied orbitals and (iii)
mutual interaction of adsorbed molecules [83,91].
The above mechanism of dissociation, i.e. first, a partial occupation of the antibon-
ding orbitals of the molecule and then bonding of the individual atoms of the molecule to
the surface, is probably operating in many other cases (e.g. N2, NO, 02). Some molecules,
CO being one example, require at least two sites to become dissociated. Two sites each
consisting of several atoms means that quite a large ensemble is required for dissociation
of such a molecule. However, some other molecules can also possibly be dissociated on
one site or perhaps even on one atom. Such is the dissociation of the H-H bond in
hydrogen or dissociation of C-H bonds in hydrocarbons.
1.2.1.3 Adsorption of molecular fragments
Not surprisingly, hydrocarbons have attracted much attention. A simple quantum
mechanical calculation has confirmed, for example, what the intuition would suggest [92]:
Structure and properties of metals and alloys 47
a CH 3 radical with one electron in a n sp 3 orbital binds to the top position, a CH 2 radical
binds to two metal atoms and a CH radical to three metal atoms on the (111) face of an
fcc metal. However, this example can immediately serve as a warning: the reality is
sometimes more complicated than the intuitive picture. Calculations involving very high
approximations have revealed that for CH 3 the optimal site is that offering the highest
coordination [93,94]. The first result seems [92] to be an artifact caused by the narrow
basis of orbitals employed.
The orbital of NH 3 which causes the binding of the molecule to the metal is the
doubly-occupied lone pair orbital of E-symmetry on the N atom. With nickel, where the
Fermi energy is low and well below the vacuum level, the highest coordination site is the
optimal one. However, with a metal such as copper, which has a lower work function, the
balance of various factors leads to a different result: the optimal site is on the top of the
copper atom [95]. The situation may be similar with the CH 3 radical, which on platinum
may be optimally bonded to a single atom [71]. However, exact theory or an experiment
must solve this problem.
Other fragments, for example, on the PF3, PF2, PF series [96] have been studied, as
has the formation of ethylidyne ( - C-CH3) [97].
The problem of dissociation of chemical bonds on the surfaces of metals, as well
as the recombination of fragments and atoms, is of great relevance for catalysis. Examples
include hydrogen dissociation [98]; C-H dissociation and recombination [98, 99], dissocia-
tion of N 2 [100] and of CO [88]. It would be extremely helpful if, in future, the theory
could also supply information on the expected behaviour of alloys.
Lennard-Jones [101] suggested a very useful picture of dissociation more than sixty
years ago. This was a general form of potential energy curves which would hold for
dissociations of simple diatomic molecules A 2. The idea is as follows. If there were no
help from the surface, the dissociation would have a potential barrier as high as the
dissociation energy D(A-A). However, if the A 2 molecule is first bound to the surface by a
weak chemisorption or by physisorption, its approach from infinity towards surface,
which leads finally to dissociation, would not encounter a barrier nearly as great as the
full D(A-A) energy. The activation energy can indeed even be zero (or < 2 kJ/mol ~) with
the mediation by weakly chemisorbed intermediates, as with H 2 on transition metals. This
situation is shown in figure 24. A similar picture should also hold for dissociation of
various other bonds, such as for example CH3-H, etc. However, recent work on molecular
adsorption dynamics has shown that, although this picture is still useful, it is a very
simplified one [15,70]. It does not account for the possibility that a molecule need not
approach the surface in a straightforward way: it can for example be repeatedly reflected
by potential walls. Furthermore, it is also difficult to express by such pictures the role of
the energy in the particular degrees of freedom and the tunnelling effects, etc.
48 chapter 1
E
. p h y s . a d s
k c h e mi s
O (AA)
D i s t a nc e , r
M e t a l - A 2
M e t a l - A
figure 24
Potential energy curves used to
A. ads describe the process of A2 disso-
ciation (Lennard-Jones):
A 2, gas (r ---) ~) ---) A2, phys.ads. ---) A2 weak chemis. ~ A, ads.atoms.
While the direct transition gas ---) Aaa~ requires a very high activation energy (D(AA)),
weakly bound forms allow a transition via a much lower barrier (zero, with weak
chemisorption).
1.2.1.4 Semi-empirical approach to the problem of chemisorption
An important initial step in the theory was taken by Eley [102], who suggested
applying to chemisorption an empirical formula due to Pauling [17]. Pauling was the first
to make an estimate of the bond strength of a molecule AB from data on A 2 and B 2
molecules. According to him the bond strength E(A-B) is an average of bond strengths in
A 2 and B2, corrected by a term containing the electronegativities x. The form of the
correction as well as the values of x were chosen to make values of E(A-B) fit a collecti-
on of experimental results. Eley, by analogy to Pauling' s approach, suggested that, for
example, for H atom chemisorption:
E(M-H) = 1/'2 { E(M-M) + E(H-H)} + 23. 06 (Xmet-XH) 2
(44)
where E(M-M) is the sublimation energy divided by the coordination number of atoms in
the metal M. For the difference in electronegativities, the value of the bond dipole of the
chemisorption bond can be taken, this being determined from measurements of work
function changes upon adsorption. For other molecules and radicals, similar equations have
been suggested [102].
Estimated bond strengths cannot be expected to be very accurate and the theory is
Structure and properties of metals and alloys 49
not straightforwardly applicable to our topic, adsorption by alloys. However, at the time it
was suggested , it was extremely important that it had been shown, that the adsorption
bond is the same kind of bond as that in free molecules.
Several other semi-empirical approaches have been suggested, but we shall confine
ourselves to mentioning only the most general one: the bond-order-conservation theory
[103]. This is based on two postulates: the f i rst postulates that the energy of a pair of
atoms forming a chemical bond depends on the distance between the nuclei r, according to
the Morse curve:
_2(r_ro) ]
E(r) Qo 2exp( - ( r - r ~ - e x p ( ~)
a a
This can be rewritten as with x substituting the exponentials
(45)
E(r) = -Qo(2X-X 2) (46)
where the exponential function x is called the bond order. For a diatomic molecule, when
r equals ro, E is equal to - Qo and the bond order is unity.
If an atom A is bound to an ensemble of several metal atoms, for example, A-M2,
the individual pair-wise bonds A-M then have a fractional order [103]. However, the
second postulate of this theory requires that the sum of the individual orders must be
unity:

x~ ( A- M i ) - 1 (47)
i=1
The main components of the theory are thus the two observables ro and Qo, which have to
be taken from experimental results, and the two foregoing postulates. With this theory it is
then possible to make predictions regarding pathways for dissociation, recombination, or
migration, for various catalytic reactions. An example of the application of this theory is
the prediction concerning the selectivity of various metals in syngas reactions [104].
The reader have probably noticed that the relation between the bond order and
bond length is formally the same as that proposed by Pauling' s equation (see eq. 28).
Thus, the same criticism which has been expressed above concerning his theory applies
here too. Further, this theory assumes that the energy of a bond is a function of only its
length, being independent of the angle which this bond makes to other pair of atoms
bonds. For the adsorption of species which make a directed bond to the surface (for
example, ,CH3, CO) and with surfaces which bind the adsorbates by spatially-oriented
orbitals, this is not a very helpful restriction. On the other hand, this theory is somewhat
flexible and suited to make the zero-approximation estimates. It has already been applied
to promoters [103] and application to alloys should be possible.
50 chapter 1
1.3 Adsorpti on of mol ecul es and radicals which are i ntermedi ates in catalytic
reactions
The understanding of results obtained in catalysis by alloys requires some
knowledge of reaction intermediates. Many have been postulated, and it is important to
realise that many are species which are very well established experimentally. Kinetic and
spectroscopic evidence exists for some of them and only the most important will be
described below. For more detailed information, the reader must consult the original
papers.
The theory concerning adsorption of H2, 02 and CO has already been briefly
discussed and we shall now concentrate our attention on other molecules. However, now
we have to rely more on experiments than on theory alone. Some aspects of the carbon
multiple pictures shown below as rt-bonding, metal-carbon multiple bonding etc., have
been analysed theoretically, but other aspects of our picture are from experiments.
1.3.1 Hydrocarbons
Hydrocarbons form a large variety of adsorbed intermediates on the surface of
metals and alloys. Burwell [ 105], who made the first inventory of species for the existence
of which good evidence was available, called this collection an organometallic zoo. The
most important species are presented in figure 25; a comment and an explanation follow.
Alkyl radicals (1) have been observed spectroscopically (for example, by EELS and
IR) but their existence, doubted ever by theoreticians and many organic chemists, was first
extablished by the very easy D2/alkane exchange [106]. The product distributions suggest
(see chapter 10) that the exchange proceeds via one of the following three species: (i)
alkyl radicals (see 1). (1) as been observed spectroscopically (for example by EELS and
IR), but their existence doubted ever by theoreticians and many organic chemists, was first
established by the very easy alkane/D2-exchange [106]. The product distribution suggest
strongly that the exchange proceeds via one of the following three species. (i) Alkyl
radicals (see (1), with these species substitution of H for D occurs in a stepwise fashion
and the product distribution has throughout a binomial character. (ii) Species multiply
bound to the surface such as 2 and 3 for which species there is now also EEL-spectral
evidence too [107-109]. (iii) Species bound as in 4 and 5 for which spectroscopic evidence
is also available [107-109]. Multiply-bound species (ii) and (iii) are responsible for
multiple exchange, i.e. the cases where more than one H atom is exchanged during one
sojourn of a molecule on the surface, in consequence of which the initial product
distribution deviates from the equilibrium (binomial) one. The deviation can be even so
pronounced that the fully exchanged alkanes appear while the products with one or two D
atoms do not desorb into the gas phase.
Structure and properties of metals and alloys 51
It is important that this chemical evidence, as just described, does exist for the
various reaction intermediates, since some sceptics doubt whether the spectroscopies
(EELS, IR) can detect reactive intermediates at all, observing instead only unreactive
spectators of the reaction.
R R
I i
CH2 CH
I II
R
I
CH
/ \
CH2 ~ CH2
-X
(b @ | 9
R CH 2
I i I i i
- - c - - c ~ c CH
Y
I i / i \ I
-X- ~ -X- -~ -~ ~ -X-
a~
@ @ | |
C \ C / C , C C
/ ,C,_-C I I
\ C / \ C / \ C / \ C C, . " " C - - C - - C - - C
/ \ / \ / , l \ '~.:t -t :: , ,
)~ )~ " ~ c c
a' [ ~ I I
| | @ @
figure 25
Various surface complexes formed by chemisorption of hydrocarbons. Only those species
for which a solid documentation exists are presented here.
Spectroscopic evidence [107,108] exists to show that species 4 is preferred by
palladium and 5 by platinum; and that both can be converted at higher temperature and
low H2 pressure into species with a higher degree of C-H bond dissociation [109-112].
The latter species (2, 6, 7) exhibit a lower reactivity in hydrogenation than those mentio-
ned above. Good evidence exists too for associatively adsorbed (horizontal oriented)
benzene (8) [113]. This all is in compliance with the theory discussed in 1.2.1.3.
52 chapter 1
For all the species 9 to 12, evidence comes only from kinetic and isotopic labelling
studies, but it is fairly solid. Carbon-labelled iso-hexanes have for example been used
[114] to show that two mechanisms operate on platinum in skeletal isomerization, one
with a C3-cyclic intermediate and one with a Cs-intermediate. Both these intermediates are
indicated in figure 26.
c 3 0 /
I~ (1)
C - C - C - C - C
{2)
5
C
l ' ,
i I I I
C - C ~ C - C - C
C - C - C
I I
C C
7 ' (
4-_
-.-t:-
C
I
C - C - C - C - C
C
i
~- C - C - C ~ C - C
figure 26
Isomerisation of a CU-labelled 2-methyl-pentane can proceed by formation of complexes in
which either 3 or 5 carbon atoms are involved. The two indicated pathways produce
differently labelled 3-methyl-pentanes.
The two distinct intermediates can thus both produce one and the same product 3-methyl-
pentane, but the products of the two pathways differ in the positions of the label. The
contributions by individual intermediates to the product distribution can be derived from
the concentration of products labelled in different places. While the arguments for the
existence of the C 3 and C 5 intermediates are strong, the technique does not tell us how the
species are bound to the surface. Indirect information suggests that species resembling 9 or
10 may be involved or an analogous species in which carbon atoms 1 and 5 are bound to
the surface. By using variously substituted pentanes, a very likely intermediate for the C 5
pathway has been shown to be species 11 [115]. Also for benzene formation a rt-
complexed C6-poly-alkene seems to be a well-documented intermediate [116,117]. Species
11 is somewhat exceptional but in c~a1313 tetrasubstituted alkanes it is probably this species
which leads to the splitting of the central C-C bond [118]. The ease with which the
Structure and properties of metals and alloys
53
variously bound species are formed decreases in the following sequence: o~B>o~T~o~8.
Metals with a high activity for hydrogenolysis of C-C bonds are all effective in
forming multiple bonds [106] and the o~B-bound species (for review see [117,119]). This
fission of bond is efficiently suppressed by diminishing the C-C particle size or by some
types of alloying. Platinum seems to be a more complicated case and here most likely the
o~ single site bound species (8) contribute to hydrogenolytic splitting [119] of hydrocar-
bons.
Figure 25 presents species for which quite solid experimental evidence already
exists. A theoretical analysis of these species comprising the prediction of their reactivity
and behaviour in the possible surface reaction is mostly missing. For example, next to
nothing can be said, on the basis of a theory, about the mechanism of the C-H or C-C
bond splitting in various species shown in figure 25. Of course, even less is known about
the predicted influence of alloying, particle size effects, etc.
1.3.2 Other molecules
Various spectroscopies (IR, EELS, XPS) and other techniques (exchange reactions,
TPD, LEED) have already yielded valuable information on surface species generated by
adsorption. As examples, we mention cyano-compounds [120], pyridine [121], pyrrole
[121], azomethane [122], trifluorophosphine [123], phenol [124] and variously substituted
benzene rings [125]. These molecules have mainly been adsorbed at lower temperatures
and, in contrast to hydrocarbons, much less is known about those species which might be
intermediates in their catalytic reactions at elevated temperatures.
Species arising from adsorption of alcohols, aldehydes, ketones and acids are
summarized in figure 27.
Alcoholate-like structures have been postulated on the basis of a very rapid
exchange of the hydroxylic H-atom [126]. The O-coordinated, di-cy coordinated and
dissociatively adsorbed species have been seen by EELS [127] as well as structures, both
mono- and bidentate, derived from acids [128]. Although details of the surface reactions
can only be speculated on, it is likely that some if not all the intermediates shown in
figure 26 participate in conversions of alcohols to aldehydes to acids, and the reverse
processes.
When overlooking this concise inventory of species, one sees the following
conclusions emerge: (i) much information exists on various intermediates, but most of
them have never been analyzed by quantum theory; (ii) most of the information relates to
metals, but very little to alloys; some will however be presented in chapter 8.
54
chapter 1
R
I
0
I
R R
I I
C C
0 0 0 0
I t \ /
* *
b i d e nt a t e
R - - C H = O R - - HC - - 0
* ~
d i - s i g ma
, e l e
O- c o o r d i na t e d bou nd
0
%
C - R
/
0
I
mo no d e nt a t e
R
!
C = 0
I
--X-
d i s s o c i a t i ve l y
a d s o r b e d
figure 27
Chemisorption complexes observed upon adsorption of oxygen-containing molecules.
1.4 Macroscopic thermodynamic theory of alloys
An important thermodynamic parameter for metals is the enthalpy of sublimation
or atomization. Figure 28 shows the values. The form of the curves can be understood in
the following terms. There are five d-orbitals and one s orbital available for bonding and
therefore the bonding is optimal when half of the valence band (see 1.1.1) is just occupied
as at tungsten. With fewer valence electrons than six the metallic cohesion is weaker than
the maxi mum possible. When there are more than six valence electrons, the antibonding
part of the valence bond is successively occupied and the cohesion is again weaker than
at the maximum. With elements that are smaller in size, the interactions between the d-d-
electrons play an increasing role and diminish the cohesion too. On the other hand, the
same phenomenon contributes to the para- and ferro- magnetism of the element. When we
pass from group 8 to group 11, not only the metallic cohesion but also the chemisorption
bond strength decreases (see 1.2). The enthalpies of adsorption of H2, CO, N2 and ethene
are very well correlated with the average metal-oxygen bond strength [129] in the highest
oxide of the metal, which in its turn is correlated with the average enthalpy of adsorption
of oxygen on the same metals.
Structure and properties of metals and alloys 55
250
O
E
t ~
, r
v
t -
O
~
N
~
E
O
O
"1"
200
150
100
50
~ 1 7 6 1 7 6 1 7 6 " ~ 1 7 6 1 7 6 1 7 6 1 7 6 1 7 6 1 7 6 ~ 1 7 6 1 7 6
.0"" "'0 ........... O.
,~176176 ~ .,, " , ,
. ' " i " " " " " " " ~ I ~ t ~ , " " 0 . . . .
~
' s ~" ~ " ' . . .
. .
s . " ~" "
, "r 0
s s - 9 ~ 1 7 6 1 7 6 1 7 6 1 7 6 1 7 6
- . ~
6
I . . . . I I 1
2 4 6 8 10
Ele m e n ts
S c . . . C u Y . ~g _ Y b . . . A u
12
f i gure 28
Heats of atomization of various transition metals.
1.4.1 A short introduction to the statistical thermodynamic description of alloys, as
random solutions
We shall confine ourselves to binary solutions and follow the literature [130] in
their description. Metallic mixtures of interest in catalysis form homogeneous liquid
solutions at sufficiently high temperatures; however, liquid alloys are in the main
uninteresting for catalysis, although some work has been done on metallic liquids, for
example, on amalgams; we have to look principally at solids formed from liquid alloys.
Several cases can be distinguished. First, a solid solution may be formed upon solidificati-
on and be homogeneous almost down to the atomic dimensions. Alternatively, clusters or
even microscrystals with distinctly different compositions may appear in the equilibrated
solid at a temperature lower than the melting point of the mixture. Let us analyse which
factors determine the phase equilibria in alloys.
56 chapter 1
Consider a solid solution of A and B with mole fractions x a and x B. We shall call
the number of nearest neighbours of a given atom s; this is twelve for fcc structure, eight
for bcc, etc. If there are a total of N atoms, then in a random solution the following pairs
have the indicated populations.
2
N . s x a N . s ( 1 - x a ) 2 (48)
A A : BB: A B : N . s ( 1 - x A ) x a
2 2
These expressions are the results of very simple statistics. To form a pair AA we need to
occupy one position (any of N) by A, the probability to find that is Nxa. From the s
neighbours of that atom are s.x a again of the A-type. Not to count the pairs twice, two
appears in the denominator. Then we ascribe to the bond between each pair a dissociation
energy -EAA, -EBB , or -EAB respectively, and if we set Etota I at zero when all atoms are
infinitively separated from each other, the total energy of our alloy is at 0 K:
Etotal - N . s [E s 4 xJ + EBB (1-XA) 2 + 2X(1-XA) E ~ ]
(49)
2
After algebraic rearrangement and after adding the Cp-term for the effect of the difference
between temperature T and absolute zero, the internal energy is:
T
+ f % aT (50)
o
There are two terms in the entropy: (i) the thermal or internal entropy So, which for pure
components is I(Cp/T)dT and (ii) the configurational entropy of mixing which is
Sco,r = - N k [x4 In x A + (1-xA) In (1-xA)] (51)
k being the Boltzman constant.
The Gibbs free energy is therefore:
G = U - T S = U - T S ~ + N k T [ x a I n x A + (1 - x A) I n (1 -;CA)]
(52)
To illustrate the results we shall analyse three cases characterized by the parameter Z: (i)
ideal solution alloys, for which
eaA + E BB ( 5 3 )
Z - 2 - 0
(ii) exothermically formed alloys (Z < 0) and (iii) the endothermically formed alloys (Z >
0). Figure 29 shows the Gibbs energy graphically, for the three indicated cases.
Structure and properties of metals and alloys 57
0 1 0
X ---->
A
figure 29
Free energy G as a function of the alloy composition x a.
-U
I
I
I
I
I
I
I
i
,,
X (1 ) X ( 2 ) 1
A A
left: ideal solutions shown and (weakly) exothermically f ormed alloys show a similar
c ur v e .
right: alloys f ormed endothermically
upper stipped line: a physical mixture of pure unalloyed elements
lower stipped line: a physical mixture of two alloys with composition XA(1) and XA(2),
respectively.
The graph indicates that formation of a mixture of two alloys ('1', ' 2' ) f r om the frozen
solution (AG") or f rom the physical mixture of elements (AG'), is thermodynamically
favoured.
- T
i s o l
I
I
I
, T cr i t
r ch
Xm
X >1
A
' B'
r i c h
figure 30
Phase diagram, showing
the limits of the (co-)
existence of phases.
An alloy with compo-
sition x m is a liquid solu-
tion at Tso l, but at T <
Tcrit it should decompose
(if the kinetics of phase
separation allows it) into
two solutions, 'A' and
'B'-rich ones.
58 chapter 1
In the ideal case (left), the form of the curve is dictated by changes in the Sconf. When the
solution is formed weakly exothermically, and no compounds are formed, the curve has a
similar form, but the contribution of enthalpy of mixing makes the curve form a deeper
valley. In the endothermic case (right), G always decreases for small concentrations of A
in B or B in A. This is due to the logarithmic form of the configurational entropy.
However, when x A is 0.5, the configurational term is small (near to zero) and the
endothermic enthalpy of mixing causes G to rise. The straight lines in the figures represent
the free energy of a physical mixture of pure components.
Let us now inspect the case at the right of figure 29 more closely. Suppose we
make a phase with a composition XA ~l~ and another one with XA ~2~. By physically mixing
these two phases, we can obtain a mixture with any composition between the indicated
ones. The free energy of these mixtures is always on the straight line connecting G1 and
G 2 . The mutual tangent connecting the points G l and G 2 at constant temperature and
pressure fulfills the condition
OG OG
or in other words, the chemical potentials are equal. This therefore represents the equilibri-
um situation. If we prepare a frozen solution of a composition x m and allow it to equilibra-
te, it will rearrange until it is converted into two phases with compositions x ~ and x ~2~.
Repeating this analysis for different temperatures, we obtain a graph in which we can
indicate the limits of existence (T, XA) of individual phases, as for example in figure 30.
When the state of an alloy approaches the limits (T, x) of solution phases,
clustering appears leading to the nuclei formation of the new phases. When the attractive
interaction between the components is strong, ordering occurs in the alloy. For example,
within the group of catalytically interesting alloys , it is observed with the Pt-Cu system.
For the ordered alloys, there are theories of different degrees of approximation and sophis-
tication [130]. The simplest is the approximation of regular solutions (Bragg-Williams
approximation). In this case U and S ~ are corrected for ordering, but the term for the
configurational entropy is kept in the form valid for ideal solution (Z = 0). This approxi-
mation has been also widely applied in dealing with the problems of surface segregation
(see chapter 4) and adsorption of gases.
Structure and properties of metals and alloys 59
With a large piece of a metal or an alloy, the surface energy, that is the energy
needed to form the surface when we cut such a piece from an infinite large block, is
negligible. However, most catalysts contain small metal particles. A frozen solution
comprising such particles might not decompose into the equilibrium phases because the
consequential energy gain could be insufficient to compensate for the energy necessary to
form new surfaces or interfaces. Some calculations illustrating this point have been
published [ 131 ]. For the catalytic chemist it is important to realize that, when a solution of
metallic elements is prepared by chemical means, it can survive equilibration and
annealing as a frozen solution, when the alloy is in the form of very small particles. It has
been seen experimentally that making the particles smaller also suppresses the ordering.
We shall also see in chapter 4 that the surface segregation is less likely to occur in small
particles.
1.4.2 Phase composition of some catalytically interesting alloys
Nickel-Copper
The early literature took for granted that this system formed a continuous series of
solutions. However, Sachtler et al. [132] showed that if this alloy is formed endothermical-
ly, it should decompose into two phases below a critical temperature Tcrit (fig.30). With
data available at that time they [132] predicted a critical temperature of about 1100K.
Another author [133] predicted the critical temperature of 450K. An experimental study
[ 134] with equilibrated co-evaporated nickel-copper films showed that the critical tempera-
ture was between 440 and 490K. Once the equilibrated solution had been formed and the
structure did not contain many defects, annealing below Tcrit did not lead to a reverse
decomposition of the solution alloy into two phases. This means that alloys used in the
early literature were probably frozen solutions when they were prepared from carbonates
by thermal decomposition and reduction, or by other procedures favouring good mixing at
elevated temperatures. On the other hand, the complicated process of slow diffusion in
solids, the necessity of demixing below Tcrit and the effects of gas induced segregation (see
chapter 4) could have been the reason why the results in the early literature were someti-
mes badly reproducible and surprisingly irregular [135].
Nickel-copper alloys also exemplify other phenomena mentioned in section 1.4.1.
For example, in one-phase solutions above Tcrit, clustering of the magnetic moments of
nickel into giant moments occurs. This has been seen by magnetic measurements [136]
and by neutron diffraction [137]. The latter study in particular offers extended information
on the kinetics of the process of nickel clustering in nickel-copper alloys and on the
ensuing steady state. For catalysis it is important to note that alloys which appear
macroscopically (e.g. by X-ray diffraction) to be homogeneous, are not really so on the
atomic scale. Thus, when comparing alloys palladium-silver and nickel-copper, one of the
60 chapter 1
differences could be the larger tendency to form clusters in the latter than in the former
system. This can also play a role in valence band electron photoemission spectra (chapters
2 and 3).
Sachtler [132] and Burton [131] have pointed out that, when alloys equilibrate into
a mixture of two phases, the phase with the lower surface energy may surround the other
when the particles are small, i.e. about 20 nm or less. Above that, the surface segregation
effects can also take place in the external shell.
Selected palladium alloys
The phase diagram of the palladium-silver alloys in figure 31 shows that at a given
temperature, the compositions of the liquid and solid phases are not exactly equal, but that
in the whole range of composition [138] solid solutions are formed. These solutions do not
deviate greatly from ideal solutions. The same holds also for palladium-gold alloys,
although in this case some authors mention a slight tendency to ordering.
The palladium-nickel system is also interesting. While the alloys mentioned above
can be used in catalysis to investigate or to exploit practically the consequences of the
dilution of an active metal (Pd) in an inactive one (Ag, Au), nickel-palladium alloys
combine two metals both of which are active, but which differ substantially in their
catalytic behaviour. For example, nickel readily forms multiple bonds with hydrocarbons
but palladium does not. Nickel is an excellent element for hydrogenolysis, but palladium is
very bad; nickel dissociates carbon monoxide at only slightly elevated temperatures, while
palladium does not do that.
The palladium-copper system does not form solid solutions in the whole range, but
only in a somewhat broad range of composition (10-100% ) [138]. Palladium forms also a
series of solutions with platinum [139].
Some platinum alloys
Platinum is one of the most important elements in catalysis. Since its physical pro-
perties can be widely manipulated by alloying with other metals, there was a hope that
this would hold also for the catalytic properties. A great deal of catalytic work has
therefore been carried out with platinum alloys.
Platinum is not an element which forms solutions so easily as nickel or palladium.
One of the exceptions is however the platinum-copper combination which forms solutions,
but with some tendency for ordering. At high temperatures continuous solid solutions are
formed but long annealing produces ordered superlattices which appear at Cu4Pt, Cu3Pt
and CuPt. The tendency to ordering is stronger than with palladium-copper, although real
intermetallic compounds are not formed [140,141 ].
The most important alloy for catalytic processes in the industry is probably
platinum-rhenium. Platinum-rhenium solutions have been prepared [142] by arc melting
Structure and properties of metals and alloys 61
and homogenization at about 2200K and at about 1300K the solubility of rhenium in
platinum was found to be about 40% wt. On the other side of the diagram, about 40% wt
of platinum is soluble in rhenium to form a hcp structure (see figure 32).
O
C
16 0 0
1400 '
1 2 0 0
1000
A t . % Pd
2 0 40 60
Li q
/
/ ,
9 6 0 . 5
Liq
I . S o ~
J i "
/
i
/
/
f
f
S o l i d
/
f
J
i
/
S o l u t i o n
8 0
(1541 ~ )
i
t
0 20 4 0 6 0 8 0 100
W t . % Pd
A g - Pd
figure 31
Phase diagram showing the (co-)existence of liguid and solid Ag-Pd solutions.
From a fundamental point of view, the platinum-gold combination is also interes-
ting. It also represents the dilution of an active by an inactive metal, whereby platinum
and gold most likely tend to form clusters in solutions. The phase diagram shown in figure
33 demonstrates that for catalysts prepared at temperatures that are interesting for
catalysts, the solution limit on the platinum rich side is low, but on the gold rich side
about 20% at of platinum dissolve in gold. This is quite useful for fundamental studies. De
Jongste et al. [145] have shown that, when frozen solutions that are not truly homogene-
ous are prepared by coreduction from solutions of precursors, the solid solutions equilibra-
te upon reduction at about 700K into a mixture of two phases with lattice contsants
(shown as upper and lower broken line in figure 33), corresponding with the phase
diagram. This is all indicated in figure 33. The triangles show the values for the initial
frozen solutions.
62 chapter 1
Ol R
kX
3 . 91 0
3 . 9 0 0
3 . 8 9 0
P! - R e
2
el I0 2 0 30 4 0 5'0 6 0 "/;0 8 0 9 0 R I
C in
, x f i gur e 32
,~o Lat t i ce cons t ant s o f s ol i d sol ut i -
4, 3o ons Pt - Re;
, , , o (Pt - f c c ; Re- hcp)
O
2 . 7 6 0
) 2 . 7 5 6
%R e
Delhez and Mittemeijer [146] have carried out a full position, width and peak-
shape analysis with the above mentioned materials and concluded that chemically-prepared
alloy particles of about 200 nm size contain a platinum-rich kernel and a gold shell.
Alloys prepared by this procedure obviously look like Sachtler' s cherries mentioned
above: a kernel and a shell both of different compositions.
Platinum-tin catalysts are used quite widely in naphtha reforming. Platinum and tin
form a series of well defined intermetallic compounds having narrow ranges for solubility
of an excess element, as can be seen in figure 34. [140,147]. It is therefore most likely
that chemically prepared mixtures of platinum and tin will tend to convert themselves
upon annealing into mixtures of various compounds, amongst which Pt 3 Sn and PtSn will
dominate.
Closing remark.
It is not our ambition to present a full treatment of an essentially metallurgical
topic discussed by this paragraph. However, we hope that with the presented information
as a base, the reader will find it easier to consult the original or reviewing literature
[138,140,148,149] when necessary. This literature contains also all information which has
been quoted throughout this chapter.
Structure and properties of metals and alloys 63
1 4 0 0 -
1 2 0 0
1 0 0 0
8 0 0
6 0 0
4 0 0
T (~
4.10
4. 05
4. 0 0
3. 95
I I I I I I I I
I
k \
\
\
\
\
\
\
\ o
\
I I I l I l
0 2 0 4'0 6'0 8'0 I O 0
AI. % S n
=C i 0 2 0 3 0 4 0 5 0 6 0 7 0 80 9 0
1800 i ~ , , L __ _~ _.._,~ jl~ _.__~ l__l~
i l I F\ J / !i \ ] i
' ~1 7 6 : i ~ t t
7 0 0 o - * ~ - - ' ~
~J I0 r JO 40 50 60 7b 80 90100
Wr . % Sn
% P t P t - S n
figure 33
Upper part: phase diagram of the Pt-Au alloy system
Lower part: alloy lattice constants of indicated composition
Full line: thermally annealed alloys
Stipped line (triangles)frozen solutions prepared by chemical co-reduction.
figure 34 (on the right)
Phase diagram for Pt-Sn alloys showing the regions of existence and coexistence of
various intermetallic compounds.
64 chapter 1
References
10
11
12
P.W.Atkins, " Physical Chemistry" , Oxford University Press (various editions)
G.M.Barrow, " Physical Chemistry",McGraw Hill (various editions)
R.S.Berry, S.A.Rice, J.Ross, "Physical Chemistry" , J.Wiley (1980)
P.A.Cox, " The Electronic Structure and Chemistry of Solids" , Oxford Science Publ.
(1987)
W.van Doom, J.Koutecky, Int.J.Quant.Chem., suppl.12 (1977) 13
J.Haydock, V.Heine, M.J.Keley, J.Phys.C, 5 (1972) 2845
G.A.Burdick, Phys.Rev. 129 (1963) 138
L.MacKinon, "Experimental Physics at Low Temperatures", Wayne State
University Press, Detroit (1966)
D.Shoenberg, Low Temp.Phys.B (1966) 680
A.R.Mackintosh, Sci.American 7 (1963) 201
P.B.T.Visscher, L.M.Falicov, Phys.Stat.Sol.(b) 54 (1972) 9
N.W.Ashcroft, N.D.Mermin, " Solid State Physics" , Holt-Saunders Int.Ed.,(1976)
p. 264
L.M.Falicov, V.Heine, Phil.Mag.Suppl. 10 (1961) 57
G.B.Scott, M.Springford, Proc.Roy.Soc.London A (1970) 115
J.A.Hoekstra, J.L.Stanford, Phys.Rev.B 8 (1973) 1416
F.M.Mueller, A.J.Freeman, J.O.Dimmock, A.M.Furdyna, Phys.Rev.B 1 (1970) 4617
F.Y.Fadin, D.D.Koeling, A.J.Freeman, T.J.Watson-Yang, Phys.Rev.B 12 (1975)
5570
J.P.Jan, A.Wenge, Low Temp.Phys.B (1973) 195
C.S.Wang, J.Callaway, Phys.Rev.B 9 (1974) 4897
W.Hume-Rothery, " The Structure of Metals and Alloys" , Inst.of Metals, Mono-
graph series (1936);
idem, " Atomic Theory for Students of Metallurgy" , Inst.of Metals, Monograph
series (1946)
W.Hume-Rothery, B.R.Coles, Adv.Phys. 3 (1954) 149
W.Hume-Rothery, Platinum Metals Rev. 10(3) (1966) 94
G.M.Schwab, Trans Faraday Soc. 42 (1946) 689
idem, Disc.Faraday Soc. 8 (1950)
N.W.Ashcroft, N.D.Mermin, " Solid State Physics" , Holt-Saunders Int.Ed.,(1976)
p.202
J.C.Slater, Phys.Rev. 92 (1953) 603
M.M.Saffren, J.C.Slater, Phys.Rev. 92 (1953) 1126
L.F.Mattheis, Phys.Rev. 134A (1964) 970
J.Koringa, Physica 13 (1947) 392
Structure and properties of metals and alloys 65
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
W.Kohn, N.Rostoker, Phys.Rev. 94 (1954) 1111
B.Segall, Phys.Rev. 124 (1961) 1797
P.Hohenberg, W.Kohn, Phys.Rev.B 136 (1964) 864
W.Kohn, L.J.Sham, Phys.Rev.A 140 (1965) 1133
N.D.Lang, Solid State Phys, Series 28 (1973) 225
S.Holloway, J.Norskov, Surf.Sci. lecture notes "Bonding at Surfaces", Liverpool
Univ.Press (1991)
J.K.Norskov, Rep.Progr.Physics 53 (1990) 1253
J.K.Norskov, S.Holloway, N.D.Lang, Surf.Sci. 137 (1984) 65
S.Holloway, J.K.Norskov, J.Chem.Soc.Faraday Trans. 83 (1987) 1935
L.Pauling, " The Nature of the Chemical Bond", Cornell Press, Ithaca, N.Y. (1960)
L.Pauling, Proc.Roy.Soc.London, A 196 (1949) 343
L.Pauling, J.Amer.Chem.Soc. 69 (1947) 542
L.Pauling, F.J.Ewing, Rev.Mod.Phys. 20 (1948) 112
L.Pauling, Phys.Rev. 54 (1938) 899
L.Pauling, Nature 161 (1948) 1019
M.Mc.D.Baker, G.I.Jenkins, Adv.Catal. 7 (1955) 1
M.Boudart, J.Amer.Chem.Soc. 72 (1950) 1040
O.Beeck, Disc.Faraday Soc. 8 (1950) 126
J.H.Sinfelt, J.L.Carter, D.J.C.Yates, J.Catal. 24 (1972) 283
W.K.Hall, P.H.Emmett, J.Phys.Chem. 63 (1959) 1102
W.Hume-Rothery, Ann.Rept.Chem.Soc.(London) 46 (1949) 42
E.Shustorovich, Surf.Sci.Rept. 61 (1) (1986) 1
N.Engel, Kem.Maandesb.(Sweden) 30(5) (1949) 53; 30(8) (1949) 97; 30(9) (1949)
105; 30(10) (1949) 114
idem, Powder Met.Bull. 7 (1954) 8
idem, Amer.Soc.Metals Trans. 57 (1964) 610
idem, Acta Met. 15 (1967) 557
L.Brewer, "Electronic Structure and Alloy Chemistry of the Transition Elements" ,
(editor: P.A.Beck) Interscience, N.Y. (1963) 221
L.Brewer, "Phase Stability in Metals and Alloys", (editors: P.Rudman, J.Stringer,
R.I.Jaffee) McGraw-Hill, N.Y. (1967) 39, 241,344, 560 (see also ref.28)
L.Brewer, Metal Trans., 4 (1973) 83
L.Brewer, Science 161 (1968)3837
J.Brewer, " Phase Stability in Metal and Alloys", (editors: P.Rudman, J.Stringer,
R.I.Jaffee) McGraw-Hill, N.Y. (1967) 241 and further
W.Hume-Rothery, Acta Met. 13 (1965) 1039
H.Zettler, G.M.Schwab, Z.physik Chem. NF 76 (1971) 38
A.R.Miedema, F.R.de Boer, P.de Chatel, J.Phys.F, Metal Physics 3 (1973) 1558
66 chapter 1
32
33
34
35
36
37
38
39
40
41
42
43
44a
b
45
A.R.Miedema, J.Less Common Met. 32 (1973) 117
J.S.Faulkner, Int.J.Quant.Chem. 5 (1971) 543
N.F.Mott, H.Jones, "The Theory of the Properties of Metals and Alloys" , Oxford
University Press (1936)
N.F.Mott, Proc.Phys.Soc.London A 62 (1949) 416
J.Friedel, Nuovo Cimento Suppl. 7 (1958) 287
J.Callaway, "Solid State Physics" , Academic Press (1958) 100
W.M.Lomer, W.Marshall, Phil.Mag. 26 (1958) 185
B.L.Mordik, Research 13 (1960) 179
D.A.Dowden, J.Chem.Soc. (1952) 242; idem Ind.Eng.Chem. 44 (1952) 977
D.D.Eley, "The Reilly lectures" , Chemistry 7 (1954) 1, Univ.of Notre Dame Press
(1954)
G.Rien~icker, Abhandlungen der Deutschen Akademi der Wissenschaften zu Berlin,
Klasse fur Geol., Biol. 3 (1955) 5 (published by Akad.Verlag, Berlin/Germany)
D.H.Seib, W.E.Spicer, Phys.Lett. 20 (1968) 1441; idem Phys.Rev. 187 (1969)
1176; idem Phys.Rev.B 2 (1970) 1676
L.Nordheim, Ann.Physik 9 (1931) 607, 641
T.Muto, Sci.Papers Inst.Phys.Chem.Res.(Tokyo) 34 (1938) 377
P.Soven, Phys.Rev. 156 (1967) 809
P.Soven, Phys.Rev. 178 (1969) 1136
S.Hufner, G.K.Wertheim, J.K.Wermick, Phys.Rev. B 8 (1973) 4511
S.Kirkpatrick, B.Velicky, H.Ehrenreich, Phys.Rev.B 1 (1970) 3250
A.Bansil, L.Schwarz, H.Ehrenreich, Phys.Rev.B 12 (1975) 2893
A.Bansil, Phys.Rev.B 20 (1979) 4025
G.M.Stocks, R.W.Williams, J.S.Faulkner, Phys.Rev.B 4 (1971) 4390
A.Z.Makzymowicz, J.Phys.F Metal Phys. 6 (1976) 1705
J.S.Faulkner, G.M.Stocks, Phys.Rev.B 21 (1980) 3222
J.Kanamori, H.Akai, N.Hamada, H.Miwa, Physica B 91 (1977) 153
L.Schwarz, H.Ehrenreich, Phys.Rev.B 6 (1972) 2223
P.Weinberger, J.Staunton, B.L.Gyorffy, J.Phys.F Metal Phys. 12 (1982)
2229
W.Temmerman, B.L.Gyorffy, G.M.Stocks, J.Phys.F Metal Phys. 8 (1978) 2461
H.Ebert, P.Weinberger, J.Voitlander, Phys.Rev.B 31 (1985) 7557
S.S.Razee, S.S.Rajput, R.Presad, A.Mokurjee, Phys.Rev.B 42 (1990) 9391
R.S.Rao, A.Bansil, H.Asonen, M.Pessa, Phys.Rev.B 29 (1984) 1713
F.Gaultier, J.van der Rest, F.Brouers, J.Phys.F Metal Phys. 5 (1975) 1975
H.W.Diehl, P.L.Leath, Phys.Rev.B 19 (1976) 587, 596
N.F.Berk, Surf.Sci. 48 (1975) 289
Structure and properties of metals and alloys
67
46
47
48
49
50
51
52
53
54
55
56
57
58
59
U.Konig, P.Marksteiner, J.Redinga, P.Weinberer, H.Ebert, Zeit.Phys.B Condensed
Matter 65 (1986) 139
D.A.Papaconstantopoulos, B.M.Klein, J.S.Faulkner, L.L.Boyer, Phys.Rev.B 18
(1978) 2784
D.E.Eastman, J.K.Cashion, A.C.Swintendick, Phys.Rev.Lett. 27 (1971) 35
S.Kirkpatrick, T.P.Eggarter, Phys.Rev.B 6 (1972) 3598
L.M.Falicov, F.Yndurain, Phys.Rev.B 12 (1975) 5664
R.C.Kittler, L.M.Falicov, J.Phys.C 9 (1976) 4259
R.C.Kittler, L.M.Falikov, Phys.Rev.B 18 (1978) 2506
S.W.Wu, J.P.Bowen, K.S.Dy, CRC Crit.Rev. Solid State Sci. 10 (1980) 43
G.Cubiotti, B.Ginatempo, J.Phys.F Metal Phys. 8 (1978) 601
J.Slechta, Phys.Stat.Sol. 120 (1983) 329
J.Slechta, J.Phys.F Metal Phys. 6 (1976) 2081
J.Haydock, " Solid State Physics" (editors: H.Ehrenreich, F.Seitz, D.Turnbull)
Academic Press, N.Y. 35 (1980) 216
F.Mejia-Lira, K.H.Bennemann, J.L.Moran-Lopez, Phys.Rev.B 26 (1982) 5398
H.L.Skriver, Phys.Rev.B 14 (1976) 5187
M.Morinaga, N.Yukawa, H.Adachi, J.Phys.Soc.Japan 53 (1984) 653
M.Morinaga, N.Youkawa, H.Adachi, J.Phys.F Metal Phys. 15 (1985) 1071
S.Pick, P.Mikusik, Chem.Phys.Lett. 189 (1992) 513
T.E.Fischer, S.R.Kelemen, K.P.Wang, K.H.Johnson, Phys.Rev.B 20 (1979) 3124
J.Kaspar, D.R.Salahub, J.Phys.F Metal Phys, 13 (1983) 311
J.Ktibler, K.H.Bennemann, R.Lapka, F.Rosel, F.Oelhafen, H.J.Guntherodt, Phys.-
Rev.B 23 (1981) 5176
R.E.Watson, J.W.Davenport, M.Weinert, Phys.Rev.B 35 (1987) 508; 36 (1987)
6396
R.E.Watson, M.Weinert, G.W.Fernando, Phys.Rev.B 43 (1991) 1446
M.O.Robbins, L.M.Falicov, Phys.Rev.B 25 (1982) 2343
J.L.Moran-Lopez, G.Kerker, K.H.Bennemann, J.Phys.F Metal Phys. 5 (1975) 1277
J.L.Moran-Lopez. G.Kerker, K.H.Bennemann, Surf.Sci. 57 (1976) 540
S.M.Foiles, M.I.Baskes, M.S.Daw, Phys.Rev.B 33 (1986) 7983
Ph.Lambin, J. P.Gaspard, J.Phys.F Metal Phys. 10 (1980) 651, 2413
M.Lundberg, Phys.Rev.B 36 (1987) 4692
S.Mukherjee, J.L.Moran-Lopez, V.Kumar, K.H.Bennemann, Phys.Rev.B 25 (1982)
730
G.Kerker, J.L.Moran-Lopez, K.H.Bennemann, Phys.Rev.B 15 (1977) 638
M.C.Desjonqueres, D.Spanjaard, Phys.Rev.B 35 (1987) 952
R.A.van Santen, "Theoretical Heterogeneous Catalysis" , World Scientific, Singapo-
re (1991)
68 chapter 1
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
N.J.Castellani, Int.J.Quant.Chem. 41 (1992) 599
J.Pancir, I.Haslingerova, P.Dobransky, Surf.Sci. 230 (1990) 255
K.Masuda, Phys.Status Solidi (b) 92 (1979) 103
H.Ueba, Phys.Status Solidi (b) 99 (1980) 763
P.Muscat, J.Phys.C Solid State Phys. 16 (1983) 3641
M.Karplus, R.N.Porter, "Atoms and Molecules", The Benjamins, USA, Menlo Park
(1970)
A.Szabo, N.S.Ostlund, "Modern Quantum Chemistry", McGraw Hill, N.Y. (1982)
I.N.Levine, "Quantum Chemistry", (3rd ed.) Allyn & Bacon (1983)
P.W.Atkins, "Molecular Quantum Mechanics", Oxford Univ.Press (2nd ed., 1983)
R.G.Parr, " The Quantum Theory of Molecular Electronic Structure",
The Benjamins (1963)
J.P.Lowe, "Quantum Chemistry", Academic Press, N.Y. (1978)
A.Rauk, " Orbital Interaction Theory of Organic Chemistry", Wiley & Sons, N.Y.
(1994)
K.Hom, A.M.Bradshaw, K.Herrmann, I.P.Batra, Solid State Commun. 31 (1979)
257
I.P.Batra, K.Herrmann, A.M.Bradshaw, K.Horn, Phys.Rev.B 20 (1979) 801
C.F.Conville, S.Somerton, D.P.Woodruff, Surf.Sci. 130 (1984) 75
J.Koutecky, Trans Faraday Soc. 54 (1958) 1038
T.B.Grimley, Proc.Phys.Soc.London 90 (1967) 751
T.B.Grimley, Proc.Phys.Soc.London 92 (1967) 776
T.B.Grimley in "Adsorption-Desorption Phenomena" (editor: T.Ricca), Academic
Press, N.Y. (1972) 215 (see also ref.75)
P.Nordlander, S.Holloway, J.K.Norskov, Surf.Sci. 136 (1984) 59
C.M.Varma, A.J.Wilson, Phys.Rev.B 22 (1980) 3795
"Quantum Chemistry Approaches to Chemisorption and Heterogeneous Catalysis"
(editor: F.Ruette), Kluwer Acad.Publ.Dordrecht (1992)
R.A.van Santen, M.C.Zonnevylle, A.P.J.Jansen, Phil.Trans.Roy.Soc.London A 341
(1992) 269
M.J.Kelly, Surf, Sci. 43 (1974) 587
J.W.Gadzuk, Surf.Sci. 43 (1974) 44
J.Koutecky, Z.Electrochem. (Bunsen Ges.) 60 (1956) 835
T.B.Grimley, Proc.Phys.Soc. 90 (1967) 751
T.L.Einstein J.R.Schrieffer, Phys.Rev.B 36 (1973) 3629
R.Schrieffer in "Dynamic Aspects of Surface Physics", Proc. E.Fermi School on
Physics, (editor: Compositori), Bologna (1974) 756
H.Hjelmberg, O.Gunnarsson, B.I.Lundquist, Surf.Sci. 68 (1977) 158
V.Matolin, M.Ribholz, N.Kruse, Surf.Sci. 245 (1992) 233
Structure and properties of metals and alloys
69
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
W.L.Jorgensen, L.Salem, " The Organic Chemist's Book of Orbitals" , Academic
Press, London (1973)
J.Blyholder, J.Phys.Chem. 68 (1964) 2772
idem J.Phys.Chem. 79 (1975) 756
F.A.Cotton, G.Wilkinson, " Advanced Inorg.Chem." J.Wiley, N.Y.(1986) 97
J.Koutecky, G.Pacchioni, P.Fantucci, Chem.Phys. 99 (1985) 87
G.Pacchioni, P.S.Bagus, J.Chem.Phys. 93 (1990) 1209
G.Pacchioni, P.S.Bagus, Phys.Rev.B 40 (1989) 6003
V.Bonacic-Koutecky, J.Koutecky, P.Fantucci, V.Ponec, J.Catal. 111 (1988) 409
N.V.Richardson, A.M.Bradshaw, Surf.Sci. 88 (1979) 255
J.N.Allison, W.A.Goddard, Surf.Sci. 115 (1982) 553
N.D.Shinn, T.E.Madey, J.Chem.Phys. 83 (1985) 5928
C.L.Allyn, T.Gustafsson, E.W.Plummer, Chem.Phys.Lett. 47 (1977) 127
W.R.Riedel, D.Menzel, Surf.Sci. 163 (1985) 39
P.M.Williams, P.Butcher, J.Wood, K.Jacobi, Phys.Rev.B 14 (1976) 325
S.R.Bare, K.Griffiths, P.Hoffmann, D.A.King, G.L.Nyberg, N.V.Richardson,
Surf.Sci. 120 (1982) 367
P.Hoffmann, S.R.Bare, D.A.King, Surf.Sci. 117 (1982) 245
C.W.Bauschlicher, Chem.Phys.Lett. 115 (1985) 535
V.Ponec in "Catalysis, Specialist Reports" 5 (1982) 48
V.Ponec, W.A.van Barneveld, IEC, Product Res.Dev. 18 (1979) 268
A.de Koster, A.P.J.Jansen, R.A.van Santen, Faraday Disc.Chem.Soc. 87 (1989) 263
J.B.Banziger, Appl.Surf.Sci. 6 (1980) 105
M.Araki, V.Ponec, J.Catal. 44 (1976) 430
R.A.van Santen, A.de Koster, Stud.Surf.Sci.& Catal. 64 (1991) 1
I.Toyoshima, G.A.Somorjai, Cat.rev.Sci.Eng. 19 (1979) 105
D.A.King, Proc. 10th Int.Congr.on Catal., Budapest, 1992, Elsevier, 1993, plenary
lecture (not printed)
H.A.C.M.Hendrickx, C.des Bouvrie, V.Ponec, J.Catal. 109 (1988) 120
P.Gelin, J.T.Yates jr., Surf.Sci. 136 (1984) L1
H.Pfnur, D.Menzel, Surf.Sci. 148 (1984) 411
M.Ban, M.A.van Hove, G.A.Somorjai, Surf.Sci. 185 (1987) 355
C.Minot, M.A.van Hove, G.A.Somorjai, Surf.Sci. 127 (1982) 441
J.Schule, P.Siegbahn, U.Wahlgren, J.Chem.Phys. 89 (1988) 6982
H.Yang, J.L.Whitten, J.Chem.Phys. 91 (1989) 126
H.Yang, J.L.Whitten, Surf.Sci. 255 (1991) 193
W.Biemolt, G.J.C.S.Kerkhof, P.J.Davies, A.P.J.Jansen, R.A.van Santen, Chem.-
Phys.Lett. 188 (1992) 477
A.W.E.Chan, R.Hoffmann, J.Chem.Phys. 92 (1990) 699
70 chapter 1
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
A.Gavezotti, M.Simonetta, Surf.Sci. 99 (1980) 453
P.E.M.Siegbahn, M.Blomberg, C.W.Bauschlicher, J.Chem.Phys. 81 (1984) 2103
P.Cremaschi, J.L.Whitten, Phys.Rev.Lett. 46 (1981) 1242
J.Y.Saillard, R.Hoffmann, J.Am.Chem.Soc. 105-6 (1984) 2006
C.Zheng, Y.Apeloig, R.Hoffmann, J.Am.Chem.Soc. 110 (1988) 749
E.Ortoleva, M.Simonetta, J.Mol.Struct. (Theochem) 149 (1987) 161
J.E.Lennard-Jones, Trans Faraday Soc. 28 (1932) 333
D.D.Eley, Disc.Faraday Soc. 8 (1950) 34
E.Shustorovich, Adv.Catal. 37 (1990) 101
E.Shustorovich, Surf.Sci.Reports 6 (1986) 1
E.Shustorovich, A.T.Bell, Surf.Sci. 253 (1991) 386
R.L.Burwell jr., J.B.Peri, Ann.Rev.Phys.Chem. 15 (1964) 131
C.Kemball, Proc.Roy.Soc.London A 207 (1951) 541
J.R.Anderson, C.Kemball, Proc.Roy.Soc.London A 223 (1954) 361
M.A.Chesters, C.de la Cruz, P.Gardner, M.McCash, P.Pudney, G.Shahid, N.Shep-
pard, J.Chem.Soc.Faraday Trans. 86 (1990) 2757
D.J.Bandy et al, Phil.Trans Roy.Soc.London A 318 (1986) 141
J.C.Bertolini, J.Massardier in " The Chemical Physics of Solid Surfaces and
Heterogeneous Catalysis", (editors: D.A.King, D.P.Woodruff) Elsevier, 1984,
Vol.3B, p. 107
L.L.Kesmodel, L.H.Dubois, G.A.Somorjai, J.Chem.Phys. 70 (1979) 2180
H.Ibach, D.L.Mills, "Electron Energy Loss Spectroscopy and Surface Vibrations",
Academic Press, N.Y. (1982)
S.B.Mohsin, M.Trenary, H.J.Robota, J.Phys.Chem. 92 (1988) 5229
G.H.Hatzikes, R.I.Masel, Surf.Sci. 185 (1983) 479
J.A.Gates, L.L.Kesmodel, Surf.Sci. 124 (1983) 68
N.R.M.Sassen, Ph.D.Thesis, Leiden University, The Netherlands (1989)
J.F.M.Aarts, N.R.M.Sassen, Surf.Sci. 214 (1989) 257
F.G.Gault, Adv.Catal. 301 (1989) 1
O.E.Finlayson, J.K.A.Clarke, J.J.Rooney, J.Chem.Soc.Faraday Trans.I, 90 (1984)
191
Z.Paal, P.Tetenyi, Raect.Kinet.Catal.Lett. 12 (1979) 131
Z.Paal, Adv.Catal. 29 (1980) 273
V.Ponec in "Chemical Physics of Solid Surfaces and Heterogeneous Catalysis"
(editors: D.A.King, D.P.Woodruff) Elsevier (1982) Vol. 1, p.365
G.Leclercq, J.Leclercq, R.Maurel, J.Catal. 50 (1979) 87
A.J.den Hartog, A.G.T.M.Bastein, V.Ponec, J.Molec.Catal. 52 (1989) 129
W.Hoffman, E.Bertel, F.P.Netzer, J.Catal. 60 (1979) 316
M.E.Kordesh, W.Steuzeland, H.Conrad, Surf.Sci. 205 (1988) 100
Structure and properties of metals and alloys 71
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
T.Szilagys, Appl.Surf.Sci. 35 (1988) 19
G.Mizutami, S.Ushioda, J.Chem.Phys. 91 (1989) 598
J.G.Searfin, C.M.Friend, J.Phys.Chem. 93 (1989) 1998
F.P.Netzer, Vacuum 41 (1990) 49
F .P.Netzer, E.Bertel, A.Goldmann, Surf.Sci. 199 (1989) 87
L.Hanley, Xingcia Guo, J.Y.Yates Jr., J.Phys.Chem. 93 (1989) 6754
M.D.Alvey, J.T.Yates Jr., J.Am.Chem.Soc. 110 (1988) 1782
Xueping Xu, C.M.Friend, J.Phys.Chem. 93 (1989) 8072
A.K.Meyers, J.B. Benzinger, Langmuir 5 (1989) 1270
J.F.M.Aarts. K.G.Phelan, Surf.Sci. 222 (1989) L853
J.R.Anderson, C.Kemball, Trans Faraday Soc. 51 (1955) 1782
A.Farkas, L.Farkas, J.Am.Chem.Soc. 61 (1939) 1336
J.L.Davis, M.A.Barteau, J.Am.Chem.Soc. 111 (1989) 1782
H.Luth, G.W.Rubloff, W.D,Grobman, Surf.Sci. 63 (1977) 325
N.R.Avery, Surf.Sci. 125 (1983) 771
N.R.Avery, A.B.Anton, B.H.Toby, W.H.Weinberg, J.Electr.Spectr.Rel.Phenom. 29
(1983) 233
M.A.Henderson, Y.Zhon, J.M.While, J.Am.Chem.Soc. 111 (1989) 1185
C.J.Houtman, M.A.Barteau, J.Catal. 130 (1991) 528
J.G.Chen, J.E.Crowell, J.T.Yates Jr., Surf.Sci. 172 (1986) 733
B.A.Sexton, Chem.Phys.Lett. 65 (1979)469
G.R.Schoofs, J.B.Benziger, Surf.Sci. 143 (1984) 359
R.J.Madix, J.L.Gland, G.E.Mitchell, Surf.Sci. 125 (1983) 481
R.P.Eischens, W.A.Pliskin, Proc. 2nd Int.Congr.on Catal., Paris, 1960, Technip.
Paris (1961) Vol. 1, p.789
K.I.Tanaka, K.Tamaru, J.Catal. 2 (1963) 366
J.C.Slater "Introduction to Chemical Physics", Dover Publ.N.Y. (1970) McGraw
Hill, 1939
D.F.Ollis, J.Catal. 23 (1971) 131
D.W.Hoffman, J.Catal. 27 (1971) 374
J.J.Burton, E.Hyman, D.G.Fedak, J.Catal. 37 (1975) 106
W.M.H.Sachtler, G.J.H.Gorgelo, J.Jongepier in " Basic Problems in Thin Films
Physics" , Proc.Int.Symp.Clausthal, 1965 (editors: R.Niedermayer, H.Mayer)
v/dHoeckx Ruprecht, G6ttingen (1966) 218
J.L.Meijering, Acta Metallurgica 5 (1957) 257
P.E.C.Franken, V.Ponec, J.Catal. 42 (1976) 398
G.K.Boreskov, Kinet.Katal. 10 (1969) 1
C.G.Robbins, H.Claus, P.A.Beck, J.Appl.Phys. 40 (1969) 2269
T.J.Hicks, B.Rainford, J.S.Kouvel, G.G.Low, Phys.Rev.Lett. 22 (1969) 531
72 chapter 1
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
P.A.Beck, Met.Trans. 2 (1971) 2015
J.S.Kouvel, J.B.Comley, Phys.Rev.Lett. 24 (1970) 598
Y.Ito, J.Akimitsu, J.Phys.Soc.Japan 35 (1973) 1000
F.Acker, R.Huquenin, Phys.Lett. 38A (1972) 343
E.Vogt, Phys.State Sol.(a) 28 (1975) 11
J.Vrijen, Ph.D.Thesis, Utrecht University, The Netherlands (1977)
J.Smithels, Metals Reference Book, Vol.II (1967) Butterworths (a collection of
phase diagrams)
J.B.Darby, K.M.Myles, Metall.Trans. 3 (1972) 653
W.B.Pearson "A Handbook of Lattice Spacings and Structures of Metals and
Alloys" , Pergamon Press, Oxford (1958)
A.Schneider, U.Esch, Z.Elektrochem. 50 (1944) 290
W.Trzebiatowski, J.Berak, Bull.Acad.Polon.Sci. C1, III, II (1954) 37
A.S.Darling, R.A. Mintern, J.C.Chaston, J.Inst.Metals 81 (1952/53) 125
L.J.van der Toorn, Ph.D.Thesis, Delft Technical University, The Netherlands
(1960)
H.de Jongste, F.J.Kuijers, V.Ponec in "Preparation of Catalysts" , (editors: B.Del-
mon, P.A.Jacobs) I (1975) 207
R.Delhez, E.J.Mittemeijer, Chem.Weekblad (1977) 105
F.M.Mittemeijer, R.Delhez, J.Appl.Phys. 49 (1978) 3875
K.Schubert, E.Jahn, Z.Metalkunde 40 (1949) 399
M.Hansen "Constitution of Binary Alloys", 2nd ed.McGraw Hill, N.Y. (1958)
J.Hafner "From Hamiltonians to Phase Diagrams" (the electronic and stat.mech.the-
ory of s,p-bonded metals and alloys), Springer, Berlin (1987)
H.Kleykamp, J.Nuclear Materials, 201 (1993) 193
J.T.Taylor, Platinum Metals Rev. 29 (1985) 74
F.R.de Boer, R.Boom, W.C.M.Mattens, A.R.Mieding, A.K.Niessen, " Cohesion in
Metals" (Trans.Metal.Alloys) North Holland Publ.Amsterdam (1988)
H.Brodowsky, H-J.Schaller, " Thermochemistry of Alloys" , Kluwer Acad.Publ.
(1989)

You might also like