Sie sind auf Seite 1von 51

CHAPTER 4 STRUCTURES OF SOLIDS

4.1

Introduction1

4.2

Model Crystal Structures.. 2


4.2.1 Bonding in Solids2
4.2.2 Atomic Packing...3
4.2.3 The Unit Cell.. 4
4.2.4 Typology of Crystal Structures.. 5

4.3

The Ionic Model. 9


4.3.1 The Radius Ratio Parameter... 9
4.3.2 Electronegativity and Crystal Structure.
10
4.3.3 Lattice Energy... 11

4.4

Crystal Structures 15
4.4.1 Metallic Elements and Alloys... 15
4.4.2 Metal Salts 17
4.4.3 Metal Oxides.22
4.4.4 Metal Hydroxo and Hydrated Compounds... 24
4.4.5 Metal Silicates...31
4.4.6 Metal Sulfides... 32

4.5

Molecular Orbital Theory and the Band Model of Solids... 33


4.5.1 Molecular Orbitals and Energy Bands.. 33
4.5.2 Metals, Insulators, and Semiconductors... 34
4.5.3 Intrinsic and Extrinsic Semiconductors... .35
4.5.4 The Fermi Level in a Solid... 36

4.6

Electronic Structures of Solids... 37


4.6.1 Molecular Orbital/Energy Band Models.. 37
4.6.2 Electronic Structures of Metal Oxides..38
4.6.3 Electronic Structures of Metal Sulfides 41

________________________________________________________________________
4.1

Introduction
In aqueous processing, our primary concern is with solution-based reactions, and

the key reagents are applied mostly in the form of dissolved aqueous species. It is
important to remember, however, that in many of the reactions of interest, we encounter
reactants and/or products in the solid phase. In this connection, it is worthy of note that
1

the majority of the elements (~ 90%) exist as solids at ordinary temperatures. In several
of the chapters that follow this one, we will be preoccupied with interfacial processes,
e.g., dissolution, crystal growth, adsorption, and ion-exchange. We shall see that the
surface properties and reactivities of solids are intimately connected with the bulk crystal
and electronic structures.
We begin this chapter by considering the types of bonding that exist in solids. We
then turn our attention to the geometrical arrangements of the bonded atoms in space.
From a consideration of the rules that govern the packing of equal-sized spheres, we
identify the basic units that constitute the building blocks for crystal structures. With the
aid of these building blocks, we review the crystal structures of representative solids.
Finally, we discuss electronic structures of solids, using molecular orbital theory as our
starting point.

4.2
4.2.1

Model Crystal Structures


Bonding in Solids
We saw in Chapters 2 and 3 that, according to molecular orbital theory, a covalent

molecule or a metal-ligand complex forms its molecular orbitals by combining the atomic
orbitals of the constituent atoms. Similarly, we can treat a solid particle as a giant
molecule consisting of an extremely large number of molecular orbitals. We recall
further from Chapters 2 and 3 that we can treat molecular orbitals of a polyatomic
molecule from two extreme points of view, i.e., the localized and delocalized bond
models. Both models are useful in the discussion of solid structures. With the localized
bond approach, we can imagine a situation where one of the bonded atoms draws the
shared electrons completely to itself. This will result in a completely ionic bond, and the
resulting giant molecule will be in reality an ionic solid, where the three-dimensional
structure is based on the electrostatic interaction (bonding) between the constituent ions.
On the other hand, if the bond is completely covalent, we get a three-dimensional
covalent solid such as diamond, where the structure consists of C-C bonds. At the other
extreme, we can consider the giant molecule from the viewpoint of the delocalized bond
model. In this multi-centered molecular orbital approach, the atomic nuclei of the solid
are viewed as embedded in a sea of electrons. This situation is characteristic of the solids
of the metallic elements, and this type of delocalized bonding has acquired the name
metallic bonding. In some cases, in addition to the ionic, covalent and metallic bonds,
other intermolecualar interactions, such as van der Waals forces and hydrogen bonding
(see Section 2.4.4), contribute to the development of the three-dimensional structure.
4.2.2

Atomic Packing
In the previous section we suggested that a solid can be usefully considered as a

giant molecule. As regards the shape of this molecule, however, we were silent. In this
section we wish to turn to this question: How are the constituent atoms of a solid
organized spatially in the solid phase? To begin, we consider the number of ways spheres
can be densely packed into space. Then we relax the constraint on dense packing; this
enables us to treat the crystal structures of ionic solids. It is helpful to organize our
thinking around two-dimensional layers of spheres in contact. Then we build three
dimensional structures by depositing these layers one on top the other, as illustrated in
Figure 4.1. We can begin the layer by layer stacking by allowing the spheres of the
second layer to fall in the cavities in the first layer. We then have two ways to add a third
layer. In one scenario, we can use an ABAB procedure, where the third layer spheres are
3

vertically colinear with first layer spheres. In the other scenario, which is based on an
ABCABC pattern, the third layer spheres are vertically colinear with the holes in the first
layer. These two structures of ABAB and ABCABC are called polytypes - they are
identical in two dimensions while they are mismatched in the third. For each of these
polytypes, the coordination number is 12.

(a)

(b)

Figure 4.1 Layer by layer build-up of three-dimensional structures (Shriver, p.110).


4.2.3

The Unit Cell


When discussing the structure of a crystalline solid, it is helpful to identify the

smallest three-dimensional unit, i.e., the unit cell, that represents the repeat unit for the
two close-packed structures discussed above. To help in visualizing the unit cell, it is
convenient to represent each atom position as a point. The resulting system of points is
termed the lattice, and by connecting the lattice points with straight lines, we can
construct the unit cell. We see from Figure 4.2a that the ABAB arrangement results in a
hexagonal unit cell; the atoms are described as hexagonally close-packed (hcp). On the
other hand, the ABCABC arrangement gives a face-centered cubic unit cell, the atoms
being called face-centered cubic (fcc).

(a)

(b)
Figure 4.2 Close-packed crystal structures: (a) Hexagonal close-packed (hcp), (b)
face-centered cubic (fcc).

(a)

(b)

(a)

(b)

Figure 4.3 Less close-packed crystal structures: (a) body-centered cubic (bcc), (b)
primitive cubic (cubic-P).
While the unit cells below involve close-packed equal-sized spheres, we can
imagine other sphere packings which are not close-packed.

A unit cell involving a

coordination number of eight is shown in Figure 4.3a (Shriver, p.112). This represents a
body-centered cubic (bcc) structure and it is characterized by a system of cube-center
5

plus cube-corner lattice points. Another non-close-packed structure, the primitive cubic
(cubic-P) is shown in Figure 4.3b. This structure is derived from the bcc structure by
removing the cube-center lattice point.

EXAMPLE 4.1 Size constraints on the occupation of interstitial holes


Consider a close-packed system of spheres of radius r. (a) Show that the largest atom that can
occupy an octahedral hole will have a radius of 0.414r. (b) Repeat (a) for a tetrahedral hole, and show that
in this case the required radius is 0.225r.
Solution

________________________________________________________________________
4.2.4

Typology of Crystal Structures


When we survey the myriad of solids available, whether the various materials are

classified as metallic elements, ionic solids, or covalent solids, we find that the various
crystal structures can be considered in terms of a small set of model crystal structures. In
these structures the coordination number ranges from twelve to two.

Metals are

generally associated with large coordination numbers (8 - 12), while molecular solids
usually have low coordination numbers (e.g., 4). For ionic solids, a coordination of 6 is
typically encountered. Four of the model structures have already been discussed above,
i.e., the face centered cubic (fcc), the hexagonal close-packed (hcp), the body centered
cubic (bcc), and simple or primitive cubic (cubic-P).
Model crystal structures for compounds with 1:1 stoichiometry (i.e., chemical
formula MX) are presented in Figure 4.4. In the rock-salt (NaCl) structure (Figure 4.4a),
chloride anions are arranged in an fcc pattern, and the sodium cations reside in the
octahedral holes. The resulting coordination is described as (6,6) since each ion has six
counterions as nearest neighbors. The cesium chloride (CsCl) structure (Figure 4.4b),
which is cubic, is derivative of the ideal bcc structure, with chloride anions at the corners
and Cs cations at the cube center.

The coordination is (8,8).

The sphalerite or

zincblende (ZnS) structure (Figure 4.4c) may be viewed in terms of an fcc array of sulfide
anions with zinc cations in one set of the tetrahedral holes. The coordination is (4,4). In
the case of the wurtzite (ZnS) structure (Figure 4.4d), the sulfur anions are arranged in an
hcp pattern and the zinc cations reside in one set of the tetrahedral holes. The nickel
arsenide (NiAs) structure (Figure 4.4e) is founded on an hcp pattern of As atoms, with
nickel atoms filling the octahedral holes. The Ni atoms are arranged into a trigonal
6

prism, within which the As atom resides.

(a) Rock salt (NaCl)

(b) Cesium chloride (CsCl)

(c) Sphalerite or zincblende (ZnS)

(d) Wurtzite (ZnS)

(e) Nickel arsenide (NiAs)


Figure 4.4 Model crystal structures for MX (1:1) compounds.

Figure 4.5 presents model structures for compounds with 1:2 (or 2:1)
stoichiometry, i.e., compounds of the type MX2 (or M2X). The rutile (titanium dioxide,
TiO2) structure (Figure 4.5a) is based on an hcp lattice of anions in which only 50% of
the octahedral holes are occcupied by the cations. The coordination number of Ti is six,
while that of oxygen is three. In the fluorite (CaF2) structure (Figure 4.5b), the Ca2+
ions are arranged as an fcc array; the fluoride ions occupy both sets of tetrahedral holes,
7

giving (8,4) coordination. For compounds with the formula M2X, the cation and anion
positions are reversed, and the resulting structure is called antifluorite. In -cristobalite
(SiO2), each Si atom is surrounded by four O atoms, while each O atom lies between two
Si atoms (Figure 4.5c). This structure may be derived from the sphalerite structure by
replacing each zinc and sulfur atom with a silicon atom and inserting oxygen atoms
between the silicon atoms. In the cadmium iodide (CdI2) structure (Figure 4.5d), a layer
of Cd atoms is sandwiched in between two layers of iodine atoms. Adjacent sandwiches
are linked only via van der Waals forces.

(a) Rutile (titanium dioxide, TiO2)

(b) Fluorite (CaF2)

(c) -Cristobalite (SiO2)

(d) Cadmium iodide (CdI2)

Figure 4.5 Model crystal structures for MX2 (M2X) compounds.

(a) ReO3

(b) Perovskite (CaTiO3)

Figure 4.6 Model crystal structures for ABX3 compounds: (a) ReO3, (b) Perovskite
(CaTiO3).

The model structure for many ABX3 compounds is the perovskite (CaTiO3)
structure. This structure, in turn, may be derived from the ReO3 structure. As can be
seen from Figure 4.6a, the ReO3 structure is cubic; Re atoms occupy cube-corners, while
O atoms are located at the edges of the unit cube. When the structure is extended beyond
the unit cell, it can be seen that it is characterized by octahedra which are linked at the
cube-corners. An important feature of the ReO3 structure is the presence of a large
twelve-coordinate central hole. By inserting a large anion A into this hole, one obtains
the perovskite structure (Figure 4.6b). In the perovskite formula of ABX3, X is typically
O2- or F-. In order to maintain electroneutrality, the total charge carried by the A and B
atoms must add up to six.
Many metal compounds of general formula AB2O4, have crystal structures that are
modeled after that of spinel (MgAl2O4). In this structure, O2- ions are organized into an
fcc pattern, and 1/8 of the tetrahedral holes are occupied by A ions, while 1/2 of the
octahedral holes contain B ions. The identity of the octahedral ions may be highlighted
by using square brackets, as in A[B2]O4. In some cases, an inverse spinel A[A,B]O4 may
form.

4.3
4.3.1

The Ionic Model


The Radius Ratio Parameter
We saw in Section 4.2.1 that, from the viewpoint of the localized bond model, the

bonding electrons in an ionic solid, MX, are assigned entirely to the more electronegative
atom (X), which then becomes the anion. The less electronegative atom (M) becomes the
cation. Thus under such circumstances, the crystal structure can be viewed as a system of
charged spheres held together by attractive electrostatic interactions. Clearly, not all
solids are completely ionic.

However, it is helpful to assume a thoroughgoing

electrostatic view of solids. Then, deviations from this ideal model provide an indication
of the relative importance of covalent and electrostatic contributions.
In covalent bonding, coordination numbers are rationalized in terms of electronic
structure, e.g., the symmetry of atomic orbitals. In the case of ionic structures, size
compatibility becomes the guiding principle. Also, it was noted above that the crystal
structure of solids are derived from close packed arrays of one kind of atom and the holes
are filled completely or partially by another kind of atom. Thus, we can consider, in the
extreme, that there will be size constraints on the size of the atom that fills the holes.
Alternatively, for a given atom, we expect size constraints on the number of larger atoms
that can surround it. A useful parameter that enables one to quantify the size constraint is
the radius ratio, r+/r-. By considering the cross-sections associated with the various lattice
structures, it can be shown that the possible coordination numbers change as summarized
in Table 4.1.

Table 4.1 Radius ratio-coordination number correlations


Coordination number
3
4
6
8

Radius ratio Range


0.155 < r+/r- < 0.23
0.23 < r+/r- < 0.414
0.414 < r+/r- < 0.732
0.732 < r+/r- < 1.00

EXAMPLE 4.2 Coordination numbers of metal salts


Calculate the relevant radius ratios and predict the coordination numbers of the ions in the
following solids: (a) NaCl (b) CsF (c) CsCl (d) ZnS.

10

Solution
(a) Referring to Table 2.4, the radius of Na + is 113 pm for CN = 4 and 116 pm for CN = 6. Based on these
radii, r+/r- = 113/167 or 116/167 = 0.695; in either case, the predicted coordination number is 6,
according to Table 4.1.
(b) For a compound such as cesium fluoride, where the anion is smaller than the cation, the radius ratio is
determined as r-/r+. Taking the relevant data from Table 2.4, r -/r+ = 119/181 = 0.658. Thus, referring
to Table 4.1, we expect a coordination number of 6.
(c) For CsCl, we have r-/r+ = 167/181 = 0.923. Accordingly, the predicted coordination number is 8.
(d)

________________________________________________________________________

Figure 4.7 A structure correlation map for MX compounds (Shriver, p.122).


4.3.2

Electronegativity and Crystal Structure


We have seen that the bonding in a compound AB is likely to be ionic when the

elctronegativity difference, , is large. At the same time, we find that ionic compounds
tend to have large coordination numbers, while covalency results in small coordination
numbers. Thus, it is not surprising that there should be a correlation between and and
the type of crystal structure. Figure 4.7 presents a structure map which illustrates the
manner in which the magnitude of the electronegativity difference is related to the
11

corresponding structure. It can be seen that when is large, the tendency is to form
compounds with relatively high coordination numbers (e.g., the rock salt and rutile
structures). On the other hand, with small values, the resulting structures tend to have
smaller coordination numbers. (See also, E. Moser and W. B. Pearson, Acta Cryst., 12,
1015 (1959)).
4.3.3

Lattice Energy (Enthalpy)


A useful parameter in assessing the stability of a given crystal structure is the

lattice energy, defined as the change in energy associated with the decomposition of a
solid into its gaseous ions:
MX (s) M+ (g) + X- (g)

HL

(4.1)

Energy is required in order to degrade the lattice, and therefore, the lattice energy is a
positive quantity. Thus, the higher the lattice energy, the greater the stability of the
corresponding crystal structure. The lattice energy also provides a means for assessing
the degree of covalency associated with the chemical bonds in the solid. If an ionic
model is assumed, then a theoretical lattice energy may be calculated and this can be
compared with the experimentally determined value.
The lattice energy can be considered as consisting of two main contributions: (1)
The electrostatic coulombic interactions between the ions (Vc), and (2) the electron
overlap repulsions (Vo).
VL = Vc + Vo

(4.2)

For two ions A and B, whose centers are separated by a distance r, the coulombic
potential energy is given by:
VAB = (zAe) (zBe)/(4orAB)

(4.3)

where zA and zB respectively represent the charges on the ions, and o is the vacuum
12

permittivity (8.85 x 10-12 C2 J-1 m-1).


To obtain the total coulombic potential energy for the crystal, the summation of
all the pair potentials is determined. The resulting expression has the form:
Vc = NA[e2/4o][zAzB/d]A

(4.4)

In this expression, NA is the Avogadro number, the second term groups a number of
fundamental constants, the third term contains parameters characteristic of the specific
ions and dimensions of the lattice (the parameter d represents the nearest-neighbor
distance), and the fourth term A is called the Madelung constant and is characteristic of
the particular crystal structure. Table 4.2 presents a selection of Madelung constants for
some model crystal structures. In general, the higher the coordination number, the higher
the Madelung constant.

Table 4.2 Madelung constants


Structure Type
Cesium chloride
Fluorite
Rock-salt
Rutile
Sphaelerite
Wurtzite

A
1.763
2.519
1.748
2.408
1.638
1.641

When two ions with closed-shell electronic structures approach each other, they
will experience mutual repulsion from the overlap of their electron clouds. The potential
energy arising from this repulsive interaction can be expressed as:
Vo = Vc + NAC' e-d/d*

(4.5)

where C' and d* are constants; it is a useful approximation to set d* = 0.345 .


In view of Equations 4.2, 4.4, and 4.5, we can write:

13

V = NA[e2/4o][zAzB/d]A + NAC' e-d/d*

(4.6)

Using the fact that the total potential energy is a minimum when dV/dd = 0, it can be
shown that:
V = (NAzAzBe2/4od)(1 - d*/d)A

(4.7)

Equation 4.7 is termed the Born-Mayer equation.


The lattice enthalpy can also be determined with the Kapustinskii equation, which
is based on the empirical observation that the actual structure of an ionic solid can be
represented by an energetically-equivalent rock-salt structure. The Kapustinskii equation
can be expressed as:
HL = -(nz-z+/d) (1 - d*/d)K

(4.8)

where d = (r+ + r-), K = 1.21 MJ mol -1, and n is the number of ions present in a formula
unit of the solid. With the aid of the Kapustinskii equation, it is possible to obtain crystal
radii, termed thermochemical radii, for complex ions. Table 4.3 presents a selection of
thermochemical radii (also see A. F. Kapustinskii, Q. Rev. Chem. Soc., 20, 203 (1956)).

Table 4.3 Thermochemical radii (Huheey, p.118).

14

________________________________________________________________________
EXAMPLE 4.3 Sphalerite versus wurtzite formation
On the basis of lattice energy considerations, determine the charge and scale parameter constraints
that lead to the preferential formation of one or the other polytypes of ZnS.
Solution
Using the subscripts s and w to signify sphalerite and wurtzite respectively, we can write that
sphalerite is preferred over wurtzite provided that Vms > Vmw, i.e.,
(zAzB)s

(d)w

(d)s

(zAzB)w

>

Aw = 1.641
As

1.638

____________________________________________________________________________________

15

4.4
4.4.1

Crystal Structures
Metallic Elements and Alloys
It was noted above (Section 4.2.1) that bonding in solids of the metallic elements

is best described in terms of delocalized bonds. As a result of the non-directional nature


of this kind of bonding, the atoms of metals in the solid state can acquire the maximum
coordination numbers allowed by radius ratio considerations. Therefore, the structures of
the metallic elements tend to be close-packed. This close packing accounts for the
typically high densities of metals. What factors determine whether a given metal prefers
the fcc or hcp structure? A useful rule of thumb is that a hexagonal structure is preferred
if the number of valence electrons/valence orbitals is low, while cubic structures are
preferred when the number is high. (Why?)
Besides hcp and fcc, other structures are also known. Thus, with increase in
temperature, it is not uncommon for metals to undergo a transition from close-packed to
less close-packed structures. The onset of structural transition can be rationalized by
recognizing that greater atomic movements are more compatible with a more loose
crystal structure. At relatively low temperatures, the tendency of a given metal to form a
close-packed structure may be related to electronic structure. If a metal is viewed in
terms of positive charges immersed in an ocean of electrons, then the higher the electron
density in this sea, the higher the atomic packing. Accordingly, the alkali metals which
have relatively low valence electrons will have a greater tendency to adopt the bcc
structure. The temperature marking the transition from close-packed to non close-packed
structures may be above room temperature (e.g., Ca, Ti, Mn), or it may be below room
temperature (e.g., Li, Na).
When two or more metals are mixed in the molten state, the resulting solid
product is termed an alloy.

At one extreme, an alloy may represent a solid solution,

where the constituent atoms are randomly intermixed. At the other extreme, the atoms in
the mixture combine to give compounds with specific compositions and structure. The
simplest alloy is a binary mixture where the majority atoms represent the solvent and the
minority atoms the solute. In substitutional solid solutions, the solute atoms are located
at sites normally occupied by the solvent atoms. In the case of interstitial solid solutions,
the solute atoms insert themselves in the spaces between the solvent atoms.

The

following set of conditions must be met in order for a substitutional solid solution to
16

form: (a) the solute atom must have a size comparable with that of the solvent atom (
15%), (b) the different metals should have similar crystal structures, and (c) the metals
must have comparable electronegativities so that the possible formation of a compound is
minimized. In the case of interstitial solid solutions, the solute atoms must have sizes
compatible with the interstitial sites in the solvent lattice. Examples of intermetallic
compounds: -brass (CuZn), MgZn2, Cu3Au, and Na5Zn21.
________________________________________________________________________
EXAMPLE 4.4. Atomic properties and alloy formation
Rationalize the following observations:
(a) Both Na and K adopt the bcc structure and many of their chemical properties are similar, and yet, they
are mutually insoluble.
(b) Of the three neighboring elements, Ni, Cu, and Zn, Cu and Ni exhibit complete mutual solubility,
whereas Cu and Zn are only partially miscible.

Solution

(a) The radius of the Na atom is 1.86 while that of the K atom is 2.26 . Thus, [(2.26 - 1.86)/2.26] =
0.18, and therefore the difference in the atomic radii exceeds the required limit of 15%. Hence the two
metals are expected to be immiscible.
(b) The metals Cu and Ni are similar with respect to electronegativities, crystal structures (fcc), and
metallic radii (Ni 1.25 , Cu 1.28 , 2.3% different). Thus, we would expect solid solutions to form.
On the other hand, even though Cu and Zn have similar electronegativities and metallic radii (Cu 1.28
, Zn 1.37 , 7 percent larger), the crystal structures are incompatible for solid solution formation (Zn
is hcp, whereas Cu is fcc).

EXAMPLE 4.5. Interstitial solid solutions


For an interstitial atom to fit snugly into an octahedral hole of a close-packed metal structure, its
radius must not exceed the hole radius of 0.414r, where r is the radius of the metal atom.
(a) Given the above consideration, determine the smallest radii of metal atoms that will be compatible

17

with H, B, C, or N as interstitial atoms.


(b) The late d-block members of the first row have radii which do not exceed 1.60 , yet these metals are
known to form interstitial solid-solutions with H, B, C, and N. How can these observations be
rationalized?
Solution
(a) We need to satisfy the condition rNM < 0.414r, the subscript NM refers to nonmetal. Thus, for r H = , rB
= , rC = , and rN = , rM > rNM/0.414.
(b) Given the condition rNM < 0.414r, B, C, and N atoms require metal atoms with radii of 1.95, 1.88, and
1.80 respectively. These radii exceed the 1.60 limit of the relevant transition metals. Thus it must
be concluded that definite compounds form, rather than interstitial solid solutions.

________________________________________________________________________
4.4.2

Metal Salts
In Chapter 3 we encountered the Lewis acid-base concept. Related to this, is the

Bronsted acid-base concept, discussed in greater detail in Chapter 5. Briefly, in the


Bronsted language, an acid is defined as a species (HA) that can donate a proton
(Equation 4.9) and a base as a species (XOH) that receives a proton (Equation 4.10). A
salt then is a compound which forms when a Bronsted acid is neutralized by a Bronsted
base:
HA = H+ + A-

(4.9)

XOH + H+ = X+ + H2O

(4.10)

XOH + HA = XA (s) + H2O

(4.11)

In this Section we consider two types of salts: (a) simple binary salts, MxAy, primarily
metal halides, and (b) complex metal salts, primarily salts of oxyanions. Hydroxysalts
(i.e., salts which contain OH- groups) and hydrated salts (i.e., salts which contain water
molecules) are discussed in Section 4.4.4.
18

Metal halides.

Two key generalizations can be made regarding the crystal

structures of metal halides: (a) For a given metal, the structure of the fluoride tends to be
different from that of the other halides. Thus fluorides tend to crystallize into threedimensional structures, whereas the chlorides, bromides and iodides prefer mostly layer
and occasionally chain structures.

(b) Fluorides and oxides with similar chemical

formulas tend to have the same crystal structures. On the other hand, Cl, Br, and I salts
tend to be isostructural with sulfides, selenides, and tellurides. Table 4.4 presents the
crystal structures for selected metal halides.

Table 4.4 The crystal structures of metal halides (Wells, p.409)


Type of structure
Infinite
3-dimensional
complexes

Formula Type
MX4
MX3
MX3
MX2
MX2

M CN:O CN
8:
6:2
8 + 1:
8:?
6:?

Name of structure
ZrF4
ReO3
YF3
Fluorite
Rutile

Examples
ZrF4
YF3
CaF2
MF2, M =

Mg, V,
MX2
MX
MX

4:2
8:?
6:6

Silica
CsCl
Sodium chloride

MX

4:4

Sphalerite

Layer structures

MX
MX2

4:4
6:?

Wurtzite
CdCl2

Layer structures

MX2

6:?

CdI2

Cr, Mn, Fe, Co,


Ni, Cu, Zn
GeO2
CsCl, CsBr, CsI
M = Li, Na, K,
Rb, CsF,

AgF,
AgCl, AgBr
CuCl, CuBr,

CuI
AgI
MCl2, M = Mg,
Mn, Fe, Co, Ni
MI2, MBr2;
M= Mg, Ti, V, Cr,

Mn, Fe, Co, TiCl2,


VCl2, CuCl2,
NiBr2,
Chain structures MX2
MX
Molecular
structures
MX2

4:
2:

PdCl2
AuI

CuBr2
PdCl2
AuI

2:

HgCl2

HgCl2
19

Complex Salts.

On the basis of the electronegativities of the atoms A and B,

compounds of the type AxByOz may be classified into two main groups, i.e., complex
oxides, and oxysalts. When A and B have similar electropositive character, the solid may
be viewed as a system of A, B, and O ions, and the compound is called a complex oxide.
On the other hand, if the crystal structure comprises a metal ion A and a complex anion
BOm, then the compound is called an oxysalt. In this case, A is electropositive, while B
is electronegative, and the compound is called a salt because it may be synthesized by
reacting to a basic oxide (A-containing compound) with an acidic oxide (B-containing
compound). In an oxysalt, B is typically a nonmetal, however, in some cases it can be a
highly charged transition metal cation. It must be noted that this division of complex oxy
compounds into oxysalts and complex oxides may be ambiguous in some cases.
Complex oxides are discussed in Section 4.5.3.

Table 4.5 presents a selection of

oxyanions with their corresponding shapes.

Table 4.5 Oxyanions and their shapes (Wells, p.512)


Non-linear
Planar
Pyramidal
Tetrahedral

BO3

3-

2-

CO3

NO2NO3-

ClO2SO32-

SiO44-

PO43AsO43VO43-

SO42-

Tetragonal pyramidal
Octahedral

TeO66-

ClO3BrO3IO3ClO4IO4MnO4IO53ReO53IO65-

When an oxyanion interacts with a metal ion, the bonding atom is the oxygen. It
is found that the cation/oxide ratio can be used as a predictor of the crystal structures of
oxysalts. This is illustrated in Table 4.6 for a group of univalent nitrates, divalent
carbonates, and tetravalent borates. (Pauling, p.547). It can be seen that the radius ratio
20

of 0.67 represents a transition between the calcite (C.N. = 6) and aragonite (C.N. = 9)
structures of calcium carbonate.

21

Table 4.6 Radius ratio-structure correlations for selected oxysalts (after Pauling, p.547)
Calcite structure

Aragonite structure

LiNO3
NaNO3

0.34
0.54

MgCO3
ZnCO3
CdCO3
CaCO3

0.47
0.50
0.65
0.67

ScBO3
InBO3
YBO3

0.60
0.59
0.68

KNO3

0.76

CaCO3
SrCO3
BaCO3

0.67
0.75
0.87

LaBO3

0.79

RbNO3 structure

RbNO3
0.84
CsNO3
0.96
___________________________________________________________________________
A useful concept in examining the stabilities of complex ionic crystals is the
electrostatic bond strength (EBS). Consider a cation of charge +ze and coordination
number n (i.e., the central metal ion is engaged in n M-O bonds). Then we obtain the
electrostatic bond strength by distributing the total cation charge among the n M-O
bonds, i.e.,
EBS = z/n

(4.12)

________________________________________________________________________
EXAMPLE 4.6 The electrostatic bond strengths of M-O bonds in MOn polyhedra.
Determine the EBS for the M-O bonds in the following polyhedra: (a) SiO4, (b) AlO6, (c) MgO6.
Solution
(a) For SiO4, z = 4, n = 4, and therefore, using Equation 4.15, EBS = z/n = 1/1 = 1.
(b) For AlO6, z = 3, n = 6, and thus, EBS = z/n = 3/6 = 1/2.
(c) For MgO6, z = 2, n = 6. Thus, EBS = z/n = 2/6 = 1/3.

________________________________________________________________________
Consider an anion XOmz-; we assume that the total anionic charge (z-) is shared
22

among the m oxygen atoms. Accordingly, the partial charge on each oxygen atom is

(-

z/m). Let us further consider a metal ion, Mz+, which forms a salt MXOm. We require
that the partial charge carried by each metal-ligand bond (i.e., the electrostatic bond
strength, EBS) be balanced locally by an equal but opposite charge on the bonded oxygen
atom; this is the principle of local charge balancing. Thus, if a given oxide anion O 2(belonging to the oxyanion XOmz-) participates in j coordination polyhedra, then we
expect that the EBS of the corresponding j M-O bonds should add up to the partial charge
on the specific oxide ion. That is,
(z/m) = EBSi = (z/n)i = j(z/n)

(4.13)

________________________________________________________________________
EXAMPLE 4.7 The electrostatic valence rule.
(a) For each of these complex anions, determine the partial charge on each oxygen atom: SiO 44-, PO43-,
SO42-, ClO4-.
(b) The following compounds adopt the calcite structure: NaNO 3, CaCO3, InBO3. Determine the number
of MOn polyhedra associated with each oxygen of the respective anions.
Solution
(a) (i)

SiO42-: For XOmz-, m=4 and z=2. Therefore, partial charge = (-z/m) = -2/4 = -1/2.

(ii)

PO43-: m=4, z=3; (-z/m) = -3/4.

(iii)

SO42-: m=4, z=2; (-z/m) = -2/4 = -1/2.

(iv)

ClO4-: m=4, z=1; (-z/m) = -1/4.

(b) The calcite structure involves octahedral coordination, i.e., MO 6. Thus, for the given metal cations, the
corresponding EBS are z/n = +1/6 (Na-O), +2/6 = +1/3 (Ca-O), +3/6 = +1/2 (In-O). The partial
charges on the anionic oxygens are: (-1/3) (NO 3-), (-2/3) (CO32-), (-3/3 = -1) (BO33-). Thus, referring to
Equation 4.13, for charge neutralization, we require:
(-1/3) (NO3-) + j[1/6 (Na-O)] = 0; i.e., j =2.

23

(-2/3) (CO32-) + j[1/3 (Ca-O)] = 0; i.e., j = 2.


(-1) (BO33-) + j[1/2 (In-O)] = 0; i.e., j = 2.
The corresponding structures are as illustrated below.

Na+
O

1/6

-1/3
NO

-2/3
CO

Ca2+
1/3

O
1/6 Na+

-1
BO

In3+
1/2

O
1/3 Ca

2+

1/2 In3+

(Also see Wells, p.327)

________________________________________________________________________
4.4.3

Metal Oxides
The earth's crust consists of ~ 90% oxygen, and inorganic materials abound with

oxygen-containing compounds. In this section, we consider (a) binary metal oxides, i.e.,
compounds that comprise oxygen and a metallic element, and (b) complex metal oxides,
i.e., compounds that contain two or more metals in combination with oxygen. The related
hydroxo compounds, hydroxides, M(OH)n, and oxyhydroxides, MO(OH), are discussed
in Section 4.4.4.
Binary Metal Oxides. Table 4.7 presents a table of the structures of selected
binary metal oxides.

A survey of the binary metal oxides leads to a number of

generalizations:
(a)

In most metal oxides the bonding is ionic.

(b)

The metal atom typically has a high coordination number, 6 or 8.

(c)

The crystal structures adopted by metal oxides tend to be the same as those
for the corresponding fluorides.

(d)

The ionic radius of O2- is typically larger than that of the corresponding
metal ion; Rb+, Cs+, and Tl+ are the exceptions.
24

(e)

The crystal structure usually consists of a close-packed array of O2- ions,


with metal ions typically in the octahedral holes.

(f)

In some oxides, the metal may be present in more than one oxidation state,
e.g., M(II), M(II), and M(IV) in M3O4.

(g) Many transition metals exhibit deviations from stoichiometry, and this results
in semiconductivity (see Sections 4.5 and 4.6).
Table 4.7 The crystal structures of binary metal oxides (Wells, p.534)
Type of structure
Infinite
3-dimensional
complexes
CeO2,

Formula Type
MO3

M CN:O CN
6:2

Name of structure
ReO3

MO2

8:4

Fluorite

Examples
WO3
ZrO2, HfO2,
ThO2, UO2

MO2

6:3

Rutile

TiO2, MoO2,

MnO2,
RuO2,PtO2
M2O3

6:4

Corundum

Al2O3, Fe2O3,
Rh2O3, Ga2O3

M2O3

6:4

A-, B-, C-M2O3

Mn2O3, Sc2O3,
Y2O3, In2O3

MO

6:6

Sodium chloride

MgO, MnO,

FeO,
CoO, NiO
MO

4:4

Wurtzite

BeO, ZnO

MO2

4:2

Silica

GeO2

M2O

2:4

Cuprite

Cu2O, Ag2O

M2O

4:8

Antifluorite

Li2O, Na2O,

K2O,
Rb2O
Layer structures

MoO3, As2O3,
PbO, SnO, Re2O7
25

Chain structures
CrO3,

HgO, SeO2,
Sb2O3

Molecular structures
RuO4, OsO4
________________________________________________________________________

26

Complex Oxides. In terms of stoichiometric formula, we can identify several


different types of complex oxides, e.g., , ABO2, ABO3, ABO4 , AB2O4. Representative
structures are presented in Table 4.8.

Table 4.8 Structures of selected complex oxides


Type

CN, A:B

Structure

Examples

ABO3

6:6

Corundum

A3+B3+O3: MVO3 (M = Fe, Ti), MnFeO3


MCrO3 (M = Fe, V, Ni)

Ilmenite

A2+B4+O3: MTiO3 (M = Fe, Co, Ni)


MMnO3 (M = Co, Ni, Mg)
MVO3 (M = Co, Ni, Mg)

12:6

Perovskite

A+B5+O3: KNbO3
A2+B4+O3: MTiO3 (M = Ca, Sr, Ba, Pb)
A3+B3+O3: LaMnO3, LaCrO3

8:4

Scheelite

A2+B6+O4: MWO4 (M = Ca, Sr, Ba, Pb)


MMoO4 (M = Ca, Sr, Ba, Pb)

6:6

Wolframite

A2+B6+O4: MWO4 (M = Mg, Fe, Co, Ni, Mn, Zn)

ABO4

AB2O4
4:6
Spinel
A2+(B3+)2O4: MAl2O4 (M = Mg, Fe, Co, Ni, Mn, Zn)
_________________________________________________________________________________
4.4.4

Metal Hydroxo and Hydrated Compounds


In this section we consider the following types of compounds:

(a) metal

hydroxides, M(OH)n, (b) metal oxyhydroxides, MO(OH), (c) basic salts, and (d)
hydrated salts. A metal oxyhydroxide may be viewed as a compound intermediate
between the corresponding oxide and hydroxide, whereas a basic salt lies in between the
salt and the hydroxide. We are concerned here with hydroxy halides and hydroxyoxysalts.
Metal Hydroxides and Oxyhydroxides.

A major factor that determines the

structures of metal hydroxides and oxyhydroxides is the manner in which the OH - ion
behaves in the vicinity of the constituent cations. Three main types of behavior may be
identified, with the OH- ion behaving as: (a) a spherically symmetric anion, (b) a
27

cylindrically symmetric (i.e., dipolar) species, and (c) a quadrupolar species. As the
cation decreases in size and increases in charge, c-type behavior becomes more favorable
in comparison with b-type behavior. In its quadrupolar form, the OH - ion can undergo
hydrogen bonding; the negative and positive regions of different OH - ions interact. As a
result of the above considerations, we may divide metal hydroxides into two main
groups, i.e., those which have no hydrogen bonds (a- and b-type behaviors), and those
with hydrogen bonds (c-type behavior). Table 4.9 presents structural information for
selected metal hydroxides and oxyhydroxides.
Table 4.9 The crystal structures of metal hydroxides and oxyhydroxides
(info from Wells, p.626 ff)
Type of structure
Infinite
complexes

Formula Type
MOH

M CN:O CN Name of structure


6:6
Sodium chloride

Examples
KOH, RbOH

M(OH)3

9:3

UCl3 (Wells, p.422)

M = La, Y, Pr,
Nd, Sm, Gd, Tb,
Dy, Er, Yb, Am

M(OH)3

6:4?

Al(OH)3

-Al(OH)3 (bayerite),
-Al(OH)3 (gibbsite)

MO.OH

7:4:3*

YO.OH

M = Y, 4f metal

MO.OH

InO.OH

-CrO.OH,
-FeO.OH

MO.OH

-MO.OH

-AlO.OH (diaspore)
-FeO.OH (goethite),
-MnO.OH, -ScO.OH,
-GaO.OH, -VO.OH

MO.OH

-MO.OH

-AlO.OH (boehmite)
-FeO.OH (lepidocrocite),
-MnO.OH manganite),
-ScO.OH

MO.OH

-CrO.OH

-CrO.OH, -CoO.OH
28

Layer structures

MOH
M(OH)2

LiOH, NaOH
Cadmium iodide

M = Mg, Ca, Mn, Fe,


Co, Ni, Cd

29

Table 4.10 OH:M ratios in hydroxy-oxysalts


OH : M
Salt
OH : M
Salt
1/5
Ca5(OH)(PO4)3
6/5
Zn5(OH)6(CO3)2
1/2
Cu2(OH)PO4
3/2
Cu2(OH)3NO3
1/1
Cu(OH)IO3
2/1
Th(OH)2SO4
_______________________________________________________________________
Hydroxysalts. The metals which tend to form basic salts include Be, Mg, Al,
several of the A subgroup transition metals (e.g., Ti, Zr), 3d metals (e.g., Fe, Co, Ni), 4f
and 5f metals (e.g., Ce, Th, U), a majority of the B subgroup elements (e.g., In, Sn, Pb,
Bi).

In nature, some important ore minerals occur as basic salts, e.g., brochantite,

Cu4(OH)6SO4, and atacamite, Cu2(OH)3Cl. Corrosion products may also occur as basic
salts, e.g., hydrozincite, Zn5(OH)6(CO3)2, forms when metallic zinc is exposed to moist
air. Other examples of basic salts include white lead, Pb3(OH)2(CO3)2, used as a pigment,
and Mg2(OH)3Cl4 H2O, a reaction product in the hydration of Sorels cement. There is a
wide variation in the OH:M ratios encountered in hydroxysalts, as illustrated in Table
4.10 for a series of hydroxy-oxysalts.

(a)

(b)

Figure 4.8 Structures of hydroxyhalides: Closed-packed spheres of (a) X(OH), (b) X(OH) 3.

30

Hydroxyhalides.
oxyhdroxides or oxides.

The structures of hydroxyfluorides tend to follow those of


Thus, for example, ZnF(OH) has the diaspore structure,

CdF(OH), the CaCl2 structure, HgF(OH), a rutile-type structure, and InF2(OH), a ReO3type structure. The most common hydroxychlorides and hydroxybromides are those of
the types, MXOH and M2X(OH)3, where M = Mg, Mn, Fe, Co, Ni, Cu(II). In the MXOH
structures, X and OH ions form a close-packed layer, as shown in Figure 4.8a (Wells,
p.160). The metal atoms reside in 25% of the octahedral sites. In the case of the
M2X(OH)3 structures, X and OH ions form a close-packed layer, as shown in Figure 4.8b
(Wells, p.160). The metal atoms reside in 50% of the octahedral sites.

H2O
H2O

OH
Al

H2O

H2O
Al
H2O

OH
H2O

4+l

H2O

H2O

8+l

OH
(H2O)4Zr
OH

Zr(H2O)4

OH
OH

OH

OH

OH
(H2O)4Zr

OH

Zr(H2O)4

Figure 4.9 Structures based on finite M-OH complexes: (a) [Al2(OH)2(H2O)8]4+,


(b) [Zr4(OH)8(H2O)16]8+
Hydroxy-oxysalts. It is helpful to think of the structures of hydroxy-oxysalts in
terms of the nature of the M-OH complex. Thus the M-OH complex or subunit may be
(a) finite, (b) infinite one-dimensional (chain), (c) infinite two-dimensional (sheet), and
(d) infinite three-dimensional. In the case of the first three situations, the oxyanions play
the important role of joining the subunits into three-dimensional structures. An example
of a structure based on a finite M-OH subunit is the basic aluminum sulfate with the
empirical formula Al2O3.2SO3.11H2O.

The structural formula can be expressed as

[Al2(OH)2(H2O)8](SO4)2.2H2O. As can be seen from Figure 4.9a, there are two OH


bridges linking two Al3+ centers, i.e., a dimeric complex. The finite M-OH subunit can
also be in the form of a ring, as exemplified by the structure of [Zr 4(OH)8(H2O)16]8+ (see
Figure 4.9b).
Structures based on M-OH chains are characteristic of many basic salts of
stoichiometry OH:M = 1. The M-OH chains are linked via oxyanions to give threedimensional structures. Figure 4.10 illustrates the Cu(OH)IO 3 structure. (Wells, Fig
31

14.16a, p.647). A number of structures based on two-dimensional M-OH subunits can be


viewed as derivatives of the CdI2 structure. First, we take the CdI2 structure and replace
Cd by M and I2 by (OH)2. Then we replace 25% of the OH groups with oxygen atoms of
oxyanions, such that the plane of the XOnz- group is perpendicular to that of the layer.

(a)
o

(b)

Cu

OH

Figure 4.10 Structures of hydroxy-oxy salts based on two-dimensional M-OH


subunits. Cu(OH)IO3: (a) Plan; (b) Elevation, with the octahedral chains
perpendicular to the plane of the paper. (Wells, p.647).

or
1
a

or

or
M+

M2+

M+
1

M2+
e

Figure 4.11 Interactions of water molecules in crystals. The larger shaded circles
represent M+, OH-, F-, or oxygen of oxy-ion (Wells, p.668).
Hydrated Salts. Here we shall be concerned with hydrates of halides, oxysalts,
32

and hydroxides. We learned in Chapter 2 that the oxygen atom in a water molecule
possesses two lone pairs of electrons. These result in two regions of negative charge on
one end of the molecule. At the same time, due to the polarization of the water molecule,
the two hydrogen atoms are relatively positive, compared with the oxygen atom. Thus
the two hydrogen atoms represent two regions of positive charge. As a result, the water
molecule may be viewed as a quadrupolar species, with a tetrahedral arrangement of the
two positive and two negative regions. The manner in which a water molecule is linked
in a hydrate may be rationalized in terms of the interactions of anions, metal ions, or
other water molecules with these charged regions, as illustrated in Figure 4.11 (Wells,
p.668).
A convenient classification, based on the coordination of the cation, is illustrated
in Table 4.11 for a hydrated salt MxAy.wH2O, where A symbolizes an anion such as Cl-,
SO42-, and n is the coordination number of the metal ion. In class AI, M is fully hydrated,
and there is excess water ((w/xn) > 1). This is a relatively small group. A representative
compound is MgCl2.12H2O. Here the cation and anion are each octahedrally surrounded
by six water molecules.

Table 4.11

w/xn
A

2
9/6
7/6
1
1
1

5/6
3/4
1/2
1/3
1/4

Classification of hydrated salts (after Wells, p.670)

M Fully Hydrated
Excess H2O (I)
No Excess H2O (II)

M Incompletely Hydrated
Excess H2O (III) Excess H2O (IV)

MgCl212H2O
FeBr29H2O
NiSO47H2O

Na2SO410H2O

Sm(BrO3)39H2O
AlCl36H2O
BeSO44H2O

CoCl26H2O
NiCl26H2O
FeCl36H2O

Na2CO310H2O

CuSO45H2O

LiClO43H2O

FeF33H2O

CaCO36H2O
CuSO43H2O
CoCl22H2O
NiCl22H2O
CaSO42H2O
33

1/6

CuSO4H2O

34

4.4.5

Metal Silicates
Silicon and oxygen make up about 75% of the earth's crust, where they exist

primarily as silicate minerals.

The basic building block in silicates is the SiO 4

tetrahedron, and as illustrated in Figure 4.12a, it is convenient to represent this as a


triangle, which symbolizes a top-down view of the tetrahedron. The complex structures
of silicates can be resolved in terms of the manner in which the SiO4 tetrahedra are
linked. Examples of the possible linkages are illustrated in Figures 4.12b-f respectively
for the disilicate anion (Si2O76-), a ring (SiO32-), an infinite single chain (SiO3)n2n-, an
infinite double chain or band, (Si4O11)n6n-, and a sheet or layer, (Si 2O5)n2n-. The charge
balance in the linked structures can be obtained by ascribing unit negative charge (-1) to a
terminal O atom and zero charge to a shared O atom. Traditional nomenclature: the label
ortho is used to signify that a species is completely hydrated, while the label meta
indicates the elimination of a mole of water from the ortho species. Table 4.12 presents a
selection of silicate compounds, with corresponding SiO4 tetrahedral linkages.

O = oxygen

= silicon

(a) SiO4 tetrahedron


(b) Disilicate anion (Si2O76-)

(c) SiO32- ring

(d) Infinite single chain ((SiO3)n2n-)

35

(e) Infinite double chain ((Si4O11)n6n-)

(f) Sheet ((Si2O5)n2n-)

Figure 4.12 SiO4 tetrahedral linkages.


Table 4.12 Silicate compounds
Building Block
Discrete orthosilicate (SiO44-) anions

Compounds
Mg2SiO4, (Mg, Fe)2SiO4 (olivine)
Na4SiO4, K4SiO4, Be2SiO4 (phenacite)
Zn2SiO4 (willemite), ZrSiO4 (zircon)
M(II)3M(III)2(SiO4)3, M(II) = Ca2+,
Mg2+, Fe2+, M(III) = Al3+, Cr3+, Fe3+ (garnets)

Discrete disilicate (Si2O76-) anions

Sc2Si2O7 (thortveitite)

Trisilicate rings (Si3O96-)

BaTiSi3O9 (benitoite),
Na2ZrSi3O9H2O (catapleite)

Hexasilicate rings (Si6O1812-)

Cu6Si6O186H2O (dioptase),
Be3Al2Si6O186H2O (beryl)

Single Chains (Pyroxenes; (SiO3)n2n-)

MgSiO3 (enstatite),
CaMg(SiO3)2 (diopside),
LiAl(SiO3)2 (spodumene)

Double Chains (Amphiboles; (Si4O11)n6n-)

Na2Fe5(OH)2(Si4O11)2 (crocidolite, blue asbestos)


(Mg,Fe)7(OH)2(Si4O11)2 (amosite)

Sheets ((Si2O5)n2n-)

Al2(OH)4Si2O5(kaolinite)
Al2(OH)2Si4O10 (pyrophyllite)
Mg3(OH)4Si2O5 (serpentine)
Mg3(OH)2Si4O10 (talc)

Al-substituted sheets ((AlSi3O10)5-)

KAl2(OH)2Si3AlO10 (muscovite, white mica)


KMg3(OH)2Si3AlO10 (phlogopite, Mg-mica)
K(Mg,Fe)3(OH)2Si3AlO10 (biotite, black mica)
36

Framework

SiO2 (quartz; tridymite, cristobalite, coesite, stishovite)


MAl2-xSi2+xO8 (feldspars)
Mx/nn+[AlxSiyO2x+2y]x- zH2O (zeolites)

37

It should be noted that silicate structures can also be considered in terms of a


close-packed array of O2- ions, where tetrahedral holes are occupied by Si4+ ions, and
octahedral and tetrahedral holes by metal ions.
4.4.6

Metal Sulfides
The electronegativity of sulfur is less than that of oxygen. As a result, M-S bonds

tend to be less polar than M-O bonds. Thus, whereas the crystal structures of metal
oxides tend to be similar to those of the corresponding fluorides, metal sulfides prefer the
structures of the more covalent chlorides, bromides, and iodides. For example, layer
structures are unusual among oxides and fluorides, but are frequently observed among
sulfides and the heavier halides. Many of the sulfides of interest in aqueous processing
have structures which are more complex than those listed in Table 4.13. However, the
structures of these more complex sulfides (Table 4.14) can often be viewed as derivatives
of the simpler structures.

Table 4.13 Crystal structures of metal sulfides (after Wells, p.749)


Type of structure
Infinite
3-dimensional
complexes

Layer structures

Chain structures

M CN:S CN

Name of structure

Examples

6:6

Sodium chloride

CaS, MnS, PbS, LaS

6:6

Nickel arsenide

FeS, CoS, NiS, VS, TiS

6:6

Pyrite

FeS2, CoS2, NiS2, MnS2, OsS2, RuS2

4:4

Sphalerite

BeS, ZnS, CdS, HgS

4:4

Wurtzite

ZnS, CdS, MnS

4:4
6:3

Cooperite
Cadmium iodide

PtS
TiS2, ZrS2, SnS2, PtS2, TaS2, HfS2

6:3

Cadmium chloride

TaS2

6:3

Molybdenum sulfide

MoS2, WS2
Sb2S3, Bi2S3, HgS

38

Table 4.14 Metal sulfides


Copper Sulfides

CuS (covellite), digenite (Cu1.8S), djurleite (Cu1.96S),


Cu2S (chalcocite), Cu1+xS (blaubleinder covellite),
Cu3AsS4 (enargite),Cu5FeS4 (bornite), CuFeS2 (chalcopyrite),
CuFe2S3 (cubanite), Cu3FeS4 (idaite)

Iron Sulfides

Fe9S10 (Im pyrrhotite (5C)), Fe11S12 (Im pyrrhotite (6C)),


Fe10S11 (Im pyrrhotite (11C)), FeS2 (pyrite), FeS2 (marcasite),
Fe1+xS (mackinawite), Fe9S10 Fe11S12 (Non-integral pyrrhotite (nC)),
FeS (troilite), Fe7S8 (monoclinic pyrrhotite (4C)), Fe3S4 (greigite),
FeAsS (arsenopyrite), FeNi2S4 (violarite), (Fe, Ni)9S11 (smythite)

Nickel Sulfides

NiS2 (vaesite), NiS (millerite), Ni3S2 (heazlewoodite), -Ni7S6,


-Ni7S6 (godlevskite), Ni3S4 (polydymite), (Ni,Fe)9S8 (pentlandite)

Cobalt Sulfides

CoS (jaipurite ?), Co3S4 (linnaeite), CoS2 (cattierite),


Co9S8 (cobalt pentlandite), (Co,Fe)AsS (cobaltite)

Silver Sulfides

-Ag2S (argentite), -Ag2S (acanthite)

Arsenic Sulfides

-AsS (realgar), -AsS (-arsenic sulfide), As2S3 (orpiment)

Zinc Sulfides

-ZnS (sphalerite), -ZnS (wurtzite)

39

4.5
4.5.1

Molecular Orbital Theory and the Band Model of Solids


Molecular Orbitals and Energy Bands
Consider the build-up of a crystal, one atom at a time. As two atoms approach

each other, a stage is reached where the higher energy orbitals (which extend furthest
away from the nucleus) begin to interact. When the distance of approach is sufficiently
close, the two outermost orbitals will combine to give two molecular orbitals. Addition
of a third atom yields a third molecular orbital, and eventually n atoms give n molecular
orbitals. Typically inorganic solid materials contain on the order of 5 x 10 22 atoms per
cm3. Thus for a particular atomic orbital, the resulting molecular orbitals generated from
the interaction of ~ 1022 atoms will be numerous and will be confined in such a limited
space that they will essentially appear as a band of closely-spaced molecular orbitals.
4.5.2

Metals, Insulators, and Semiconductors


If a band is derived from s orbitals, then the resulting band is called an s band.

Similarly, p and d bands may be constructed from p and d orbitals respectively. Just as s,
p, and d orbitals lie at different energy levels, so are s, p, and d bands typically separated
by energy gaps.

The bands are filled with electrons, starting from the lowest energy

band. The occupied bonding orbitals constitute the valence band, and the empty orbitals
represent the conduction band.

The energy of the lowest molecular orbital of the

conduction band is denoted as Ec while the upper-most energy level of the valence band
is denoted as Ev.
In general, the valence and conduction bands do not overlap, but are separated by
a band gap, Eg. The value of Eg determines whether a material is a metal, an insulator,
or a semiconductor, as illustrated in Figure 4.13.

Metals are characterized by the

presence of energy bands that are partially filled, or by occupied energy levels that
overlap with unfilled or partially occupied bands. On the other hand insulators and
semiconductors are characterized by the presence of energy gaps; E g > 2eV for an
insulator, and Eg < 2eV for a semiconductor. Table 4.15 presents values of the band gaps
for selected materials.

40

CONDUCTION
BAND

CONDUCTION
BAND

CONDUCTION
BAND

ENERGY GAP

ENERGY GAP
VALANCE BAND
VALANCE BAND

(a)

VALANCE BAND

(b)

(c)

Figure 4.13 Band model of solids: (a) Metal; (b) Semiconductor; (c) Insulator
Table 4.15 Bandgaps of Selected Materials*
Semiconductor

Type of
Conductivity
n, p
n
n, p
n
p
p
n, p
n
n

Bandgap
(eV)
0.66
1.1
2.8
2.2
2.2
1.7
0.1
2.2
0.25

Semiconductor

Type of
Conductivity
n
n
n
p
p
p
n, p
n, p
n, p

Bandgap
(eV)
~ 1.8
1.2
2.4
~1
~1
?
0.9
0.3 - 1
0.37

Ge
Ag2S
Si
Bi2S3
Bi2O3
CdS
CdO
Cu2S
Cu2O
CuS
CuO
Fe1-xS
Fe3O4
FeS2
Fe2O3
MoS2
MnO2
PbS
(Pyrolusite)
SnO2
n
3.7
Sb2S3
1.7
TiO2
n
3.0
ZnS
n
3.6 3.9
V2O5
m
2.23
CuFeS2
n
0.6
WO3
n
2.7
Cu5FeS2
p
~1
ZnO
n
3.2
FeAsS
n, p
~0.2
ZrO2
n
5.0
GaAs
n, p
1.35 1.43
PbO
n
2.8
CdTe
n, p
1.5
Nb2O5
n
3.4
CaP
n, p
2.24
Ta2O5
n
4.0
InP
p
1.3
BaTiO3
n
3.3
CdSe
n
1.74
FeTiO3
n
2.8
ZnTe
p
2.3
KTiO3
n
3.5
SiC
p
3.0
SrTiO3
n
3.4
*Based on R.T. Shuey, Semiconducting Ore Minerals, Elsevier, 1975; A. J. Nozik,
Photoelectrochemistry: Applications to Solar Energy Conversion, Ann. Rev. Phys. Chem.,
29, 189-222 (1978); K. Rajeshwar, P. Singh, and J. Dubow, Energy Conversion in
41

Photoelectrochemical Systems A Review, Electrochim. Acta, 23, 1117-1144 (1978).


4.5.3 Intrinsic and Extrinsic Semiconductors
When the magnitude of Eg is sufficiently low, it is possible for thermal agitation to
promote an electron from the valence band into the conduction band of a pure material,
thereby producing an electron hole in the valence band. This material is then termed an
intrinsic semiconductor. A material can also acquire conduction band electrons and
valence band holes by receiving impurities. Consider the addition of phosphorus, which
has five valence electrons, to germanium, which has four valence electrons. In order for
P to fit into the Ge structure, only four valence electrons are required. The extra electron
from P remains under the influence of the +5 charged P nucleus and stays in an energy
level which is within the Ge band gap. If the electron is excited out of this localized band
into the Ge conduction band, then the P becomes positively charged. At the same time
the Ge lattice becomes negatively charged and is therefore called an n-type
semiconductor. Phosphorus is termed a donor impurity since it donates electrons to the
Ge conduction band.
On the other hand, when a boron atom, with three valence electrons, is introduced
in the germanium crystal structure, it is necessary to find a fourth valence electron in
order to fit this impurity into the Ge structure. The needed electron is taken from the Ge
valence band, thereby creating a hole with an effective positive charge. The B-doped
material is called a p-type semiconductor since B has donated a positive charge to the Ge
energy-band structure.
4.5.4

The Fermi Level in a Solid


Consider a solid constructed by appropriating an s electron from each of N atoms.

At absolute temperature (T = 0) one-half of the lower-lying orbitals will be occupied.


The topmost occupied orbital is designated the Fermi level and it is located midway
between the band. At higher temperatures, the probability (f) that an energy level with
energy E, is occupied by an electron is given by the Fermi distribution function:
f = 1/[1 + exp(E-EF)/kT]

(4.14)

where EF is termed the Fermi energy, and k is the Boltzmann constant. The Fermi energy
is temperature dependent; at T = 0 it is simply the energy associated with the Fermi level.
42

When E = EF, f = 1/2. That is, the probability of occupancy is one-half when a level is at
the Fermi energy.
The position of the Fermi level is dependent on the nature of the material. For a
metal, the energies of both the occupied and vacant states are located near E F. At room
temperature, the Fermi level of an intrinsic semiconductor lies in the center of the band
gap.

On the other hand for extrinsic semiconductors, E F occurs just beneath the

conduction band in n-type materials, and just above the valence band in p-type materials.
When the energy level Ec lies significantly above the Fermi level (i.e., (Ec - EF) >
2kT), then the probability that it is occupied is given by Equation 4.14 as:
f = exp-(Ec - EF)/kT

(4.15)

Thus, if the conduction band is characterized by a density of energy levels (i.e., number
of energy levels per unit volume) of Nc, then the electron density (n) in the conduction
band is given by
n = Ncexp[-(Ec - EF)/kT]

(4.16)

Similarly if (EF - Ev) > 2kT, then the density of holes (p) in the valence band is given by:
p = Nvexp[-(EF - Ev)/kT]

(4.17)

where Nv is the density of energy levels in the valence band. It follows from Equations
4.16 and 4.17 that:
np = NcNvexp[-(Ec-Ev)/kT] = Nc Nv exp(-Eg/kT)

(4.18)

43

4.6

Electronic Structures of Solids

4.6.1

Molecular Orbital/Energy Band Models


In light of the crystal structural characteristics reviewed above, the basic building

blocks of oxides such as SiO2, TiO2, and NiO can be taken as the SiO 4, TiO6, and NiO6
polyhedra respectively. Similarly, in the case of sulfides, representative building blocks
are ZnS4 (for ZnS), CuS4, and FeS4 (for CuFeS2), and FeS6 for FeS2. The formation of
these clusters can be visualized in terms of the interaction between the outer atomic
orbitals associated with the metal and oxygen or sulfur atoms. Molecular orbital energy
level diagrams for the clusters can be derived from quantum chemical calculations.
Figure 4.14 shows the results for (ZnS 4)6- and (CuS4)6-; the energy scale is taken relative
to the non-bonding S 3p type orbitals (i.e., 1t 1 for the tetrahedral clusters). In solid
materials, the molecular orbitals of the numerous clusters overlap to give bands, and this
is illustrated by the corresponding band diagram for sphalerite. In terms of the band
model, the highest occupied molecular orbital (HOMO) represents the topmost layer of
the valence band, whereas the lowest unoccupied molecular orbital (LUMO) represents
the lowest energy level in the conduction band.

5
4

2+

ZnS4 (Zn)

CuS4

3a 1

Orbital Energy (eV)

2
4t2

1
0
-1
-2
-3
-4
-5

2e
lt 1

4t 2
1t 1
2e

3t 2
2a 1

3t2
1e 2t2
,
2a1

1e
2t 2
Sphalerite

Antibonding M-S Conduction Band


Crystal-Field-Type Levels
Nonbonding Sulfur Band
Bonding M-S Band

Figure 4.14 Energy level diagrams showing the electronic structures of the (ZnS 4)6and (CuS4)6- clusters, and the band model of sphalerite (after Vaughan and Tossell).

44

4.6.2

Electronic Structures of Metal Oxides


On the basis of the origin of their energy gaps, transition metal oxides may be

classified into empty d-shell (d0, non-d-band?) oxides, closed d-shell (d-band?) oxides,
and open d-shell (d-band?) oxides, as indicated in Tables 4.16, 4.17, and 4.18 for selected
compounds. The metal oxides that fall under the d 0 category come from the earlier
transition elements: up to Mn, Ru, and Os respectively for the 3d, 4d, and 5d series.
Among the transition metal oxides, these materials exhibit the simplest electronic
structures. To illustrate the general characteristics of the electronic structure of d 0 metal
oxides, we turn to rutile TiO2.

We know from the crystal structure that Ti4+ is

octahedrally coordinated to O2-.

Molecular orbital calculations based on the TiO68-

cluster give the energy levels shown in Figure 4.15. We can identify five groups of
molecular orbitals: (1) the O 2s nonbonding orbitals (5a 1g, 4t1u, 1eg), (2) the bonding
orbitals (5t1u, 6a1g, 1t2g, 2eg), (3) the O 2p nonbonding orbitals (1t2u, 6t1u, 1t1g), (4) the
crystal-field orbitals (2t2g, 3eg) and (5) the anti-bonding conduction band orbitals (7a 1g,
7t1u). The main feature we want to note here is the nature of the highest filled orbital and
the lowest unfilled orbital, since these represent Ev and Ec respectively. The filled orbital
with the highest energy is the 1t 1g, and this represents the top of the valence band. On the
other hand, the 2t2g is the lowest-lying unoccupied molecular orbital.
Table 4.16 Band gaps for selected d0 metal oxides (After Cox, 1992)
Binary Compounds
3d0

Compound
TiO2 (rutile)
TiO2 (anatase)

Eg (eV)
3.0
3.2

V2O5
CrO3

2.2
2

Nb2O5

3.9

MoO3
Ta2O5

3.0
4.2

WO3

2.6

4d0

5d0

Ternary Compounds
Compound
MgTiO3
SrTiO3
BaTiO3
La2Ti2O7

Eg (eV)
3.7
3.4
3.2
4.0

SrZrO3
LiNbO3
KNbO3

5.4
3.8
3.3

LiTaO3
NaTiO3

3.8
3.8

45

Table 4.17 Electronic properties of selected closed-shell dn metal oxides (After Cox, 1992)
Compound Ionic Formula
Cu2O
(Cu+)2O2AgO
Ag+Ag3+(O2-)2
PdO
LaCoO3
LaRhO3

Pd2+O2La3+Co3+(O2-)3
La3+Rh3+(O2-)3

Elec. Config. Metal Coordination


3d10
2 (Linear)
10
8
4d , 4d
2 (Linear), 4 (Sq.
plan.)
8
4d
4 (Square planar)
3d6
6 (Octahedral)
6
4d
6 (Octahedral)

Eg (eV)
2.16
1
0.1
1.6

Table 4.18 Electronic properties of selected open-shell dn metal oxides (After Cox,
1992)
Binary Compounds
Compound
Spin
Eg (eV)
3d0
Cr2O3
MnO
MnO2
FeO
-Fe2O3
CoO
NiO
CuO

3/2
5/2
3/2
2
5/2
3/2
1
1/2

3.3
3.6

2.4
1.9
2.6
3.8
1.4

Ternary Compounds
Compound
Spin
LaVO3
1
LaCrO3
3/2
LaMnO3
2
LaFeO3
Y3Fe5O12

5/2
5/2

Eg (eV)

Closed d-shell metal oxides. The characteristic feature of this category of metal
oxides is the presence of a closed shell of d-electrons and the fact that the uppermost
orbital of the valence band has d-character. In the case of the d 10 compound, Cu2O, the
closed shell arises because all the five d orbitals are fully occupied, with an electron-pair
in each orbital. The band gap is bounded by the lower-lying filled Cu 3d levels, and the
higher-lying but empty Cu 4s level.

The chemical formula, AgO, easily gives the

impression that this compound contains silver in the plus two oxidation state (Ag(II)). In
fact, this is a mixed valency compound, whose ionic formula involves Ag(I), d 10, and
Ag(II), d8. As found in the case of Cu(I) above, in the Ag(I) oxidation state, the silver
atom is coordinated to two oxygen atoms. Also, as with Cu +, Ag+ is closed-shelled since
all available d orbitals are filled. The Ag(II) ion is tetrahedrally coordinated to four
oxygens. The band gap probably lies between the filled Ag+ 4d orbitals and the empty
Ag3+ 4d orbitals.
46

Figure 4.15 Molecular orbital diagram for the TiO68- cluster (Tossell and Vaughan,
p.188).
Open-shell dn metal oxides. The oxides that fall under this category contain
unpaired electrons and therefore they have magnetic properties.
4.6.3

Electronic Structure of Metal Sulfides


Molecular orbital energy level diagrams for the tetrahedral clusters of Zn 2+, Cu+,

Cu2+, Fe2+, and Fe3+, and the octahedral cluster of Fe2+, have been derived from quantum
chemical calculations by Vaughan and Tossell. The results are presented in Figure 4.16.
In comparing the different clusters, the following observations can be made:

47

2+

FeS4 (Fe)

5
4

CuS4 (Cu)

2+
ZnS6
(Zn)
4

+3

FeS4 (Fe)

3a1

3a 1
4t2
e

t2
e

Orbital Energy (eV)

2
1
0
-1

4t 2
1t 1
2e
3t 2

1t 1
3t 2
1e
2t 2
2a 1

t2
e

-2
-3
-4
-5

2a 1

t2

4t 2

2e

1e
2t 2
Sphalerite

Ferroan Sphalerite

Chalcopyrite

Antibonding M-S Conduction Band

Nonbonding Sulfer Band

Crystal-Field-Type Levels

Bonding M-S Band

Figure 4.16 Energy level diagrams showing the electronic structures of metal sulfide
clusters (after Vaughan and Tossell).
(1)

The highest occupied molecular orbital (HOMO) of the ZnS4 cluster (the 4t2
orbital) is of predominantly nonbonding sulfur 3p character. In contrast, the
HOMOs of the other clusters are represented by antibonding metal 3d-type
orbitals. The HOMOs of (CuS4)7- and (CuS4)6- are filled and partially filled
4t2 orbitals respectively.

In the case of (FeS 4)6-, the four unpaired 3d

electrons in this structure cause the energy levels to divide into spin-up and
spin-down groups; the empty 10t2 and partially occupied 3e molecular
orbitals lie at the top of the valence band. The 2t 2g orbital serves as the
HOMO and valence band for (FeS6)10-.
(2)

For both (ZnS4)6- and (CuS4)7-, the LUMO is the antibonding 3a1 orbital.
However, since the 4t2 orbital of the (CuS4)6- cluster is incompletely filled, it
represents the LUMO. In the case of (FeS 4)6-, the empty antibonding 9a1
(mostly Fe 4s) and 11t2 (mostly S 3p) orbitals represent the conduction
band, while the empty 3eg orbitals serve as the LUMO for the cluster, and
therefore, for FeS2.

48

(3)

In the (ZnS4)6- electronic structure, the metal 3d-type orbitals are buried far
below the bonding M-S orbitals. This is in contrast to the other clusters,
where the opposite ordering is observed.
+
CuS7
4 (Cu)
5
4
Orbital Energy (eV)

3
2

3+
FeS5(Fe)
4

3a1
4t 2
e

t
e2

1
0
-1
-2

t2

1t1
3t2
1e
2t2
2a1

4t 2
2e

-3
-4
Chalcopyrite

-5

Antibonding M-S
Conduction Band

Nonbonding
Sulfur Band

Crystal-FieldType Levels

Bonding
M-S Band

Figure 4.17 The electronic structure of chalcopyrite CuFeS2 (after Vaughan and
Tossell).

49

D-BAND SEMICONDUCTOR
6
5
4
3

Energy (eV)

2
1
0
-1
-2
-3
-4
-5
-6
-7

eg
t 2g

NON-D-BAND SEMICOND.

5
Crystal-FieldType Levels

t 1g
t 1u

Antibonding M-S
Conduction Band

3
2
1
0

eg
t 2u

-1

t 1u
t 2g
a1g

-3

Bonding M-S
Band

3a1

-2
-4

4t 2
1t 1
2e

Nonbonding
Sulfur Band

3t 2

Bonding M-S
Band

2a1
1e
2t 2

Crystal-FieldType Levels

-5

Pyrite

Sphalerite

(FeS 2)

(ZnS)

Figure 4.18 Comparison of the electronic structures of pyrite (FeS2) and sphalerite
(ZnS) (after Vaughan and Tossell)

50

Further Reading
1.

A. F. Wells, Structural Inorganic Chemistry, 5th ed., Oxford, New York, 1984.

2.

P. A. Cox, Transition Metal Oxides, Oxford, New York, 1992.

3.

P. A. Cox, The Electronic Structure and Chemistry of Solids, Oxford, New York,
1987.

4.

D. J. Vaughan and J. R. Craig, Mineral Chemistry of Metal Sulfides, Cambridge,


New York, 1978.

5.

J. A. Tossell and D. J. Vaughan, Theoretical Geochemistry: Application of Quantum


Mechanics in the Earth and Mineral Sciences, Oxford, New York, 1992.

6.

L. Pauling, The Nature of the Chemical Bond, 3rd. ed., Cornell University Press,
Ithaca, New York, 1960.

7.

J. E. Huheey, E.A. Keiter, and R. L. Keiter, Inorganic Chemistry. Principles of


Structure and Reactivity, 4th ed., Harper Colins, 1993.

8.

A. G. Sharpe, Inorganic Chemistry, 3rd. ed., Longman, London, 1992.

9.

D. E. Shriver, P. W. Atkins, and C. H. Langford, Inorganic Chemistry, Freeman,


New York, 1990.

10.

B. Webster, Chemical Bonding Theory, Blackwell, Oxford, 1990.

11.

F. A. Cotton and G. Wilkinson, Advanced Inorganic Chemistry, 5th ed., Wiley, New
York, 1988.

12.

L. E. Orgel, An Introduction to Transition - Metal Chemistry: Ligand-Field Theory,


2nd ed., Wiley, New York, 1966.

13.

F. Liebau, Structural Chemistry of Silicates, Springer-Verlag, New York, 1985.

51

Das könnte Ihnen auch gefallen