SlideShare a Scribd company logo
1 of 48
Download to read offline
between clay and nonclay minerals is that the nonclays 
are composed primarily of bulky particles; whereas, 
the particles of most of the clay minerals are platy, and 
in a few cases they are needle shaped or tubular. 
The great range in soil particle sizes in relation to 
other particulate materials, electromagnetic wave 
lengths, and other size-dependent factors can be seen 
in Fig. 3.2. The liquid phase of most soil systems is 
composed of water containing various types and 
amounts of dissolved electrolytes. Organic compounds, 
both soluble and immiscible, are found in soils at sites 
Copyrighted Material 
35 
CHAPTER 3 
Soil Mineralogy 
Figure 3.1 Particle size ranges in soils. 
3.1 IMPORTANCE OF SOIL MINERALOGY IN 
GEOTECHNICAL ENGINEERING 
Soil is composed of solid particles, liquid, and gas and 
ranges from very soft, organic deposits through less 
compressible clays and sands to soft rock. The solid 
particles vary in size from large boulders to minute 
particles that are visible only with the aid of the elec-tron 
microscope. Particle shapes range from nearly 
spherical, bulky grains to thin, flat plates and long, 
slender needles. Some organic material and noncrys-talline 
inorganic components are found in most natural 
fine-grained soils. A soil may contain virtually any el-ement 
contained in Earth’s crust; however, by far the 
most abundant are oxygen, silicon, hydrogen, and alu-minum. 
These elements, along with calcium, sodium, 
potassium, magnesium, and carbon, comprise over 99 
percent of the solid mass of soils worldwide. Atoms 
of these elements are organized into various crystalline 
forms to yield the common minerals found in soil. 
Crystalline minerals comprise the greatest proportion 
of most soils encountered in engineering practice, and 
the amount of nonclay material usually exceeds the 
amount of clay. Nonetheless, clay and organic matter 
in a soil usually influence properties in a manner far 
greater than their abundance. 
Mineralogy is the primary factor controlling the size, 
shape, and properties of soil particles. These same fac-tors 
determine the possible ranges of physical and 
chemical properties of any given soil; therefore, a 
priori knowledge of what minerals are in a soil pro-vides 
intuitive insight as to its behavior. Commonly 
defined particle size ranges are shown in Fig. 3.1. The 
divisions between gravel, sand, silt, and clay sizes are 
arbitrary but convenient. Particles smaller than about 
200 mesh sieve size (0.074 mm), which is the bound-ary 
between sand and silt sizes, cannot be seen by the 
naked eye. Clay can refer both to a size and to a class 
of minerals. As a size term, it refers to all constituents 
of a soil smaller than a particular size, usually 0.002 
mm (2 m) in engineering classifications. As a mineral 
term, it refers to specific clay minerals that are distin-guished 
by (1) small particle size, (2) a net negative 
electrical charge, (3) plasticity when mixed with water, 
and (4) high weathering resistance. Clay minerals are 
primarily hydrous aluminum silicates. Not all clay par-ticles 
are smaller than 2 m, and not all nonclay par-ticles 
are coarser than 2 m; however, the amount of 
clay mineral in a soil is often closely approximated by 
the amount of material finer than 2 m. Thus, it is 
useful to use the terms clay size and clay mineral con-tent 
to avoid confusion. A further important difference 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
36 
Copyrighted Material 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
37 
Copyrighted Material 
Figure 3.2 Characteristics of particles and particle dispersoids (adapted from Stanford Re-search 
Institute Journal, Third Quarter, 1961). 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
38 3 SOIL MINERALOGY 
that have been affected by chemical spills, leaking 
wastes, and contaminated groundwater. The gas phase, 
in partially saturated soils, is usually air, although or-ganic 
gases may be present in zones of high biological 
activity or in chemically contaminated soils. 
The mechanical properties of soils depend directly 
on interactions of these phases with each other and 
with applied potentials (e.g., stress, hydraulic head, 
electrical potential, and temperature). Because of these 
interactions, we cannot understand soil behavior in 
terms of the solid particles alone. Nonetheless, the 
structure of these particles tells us a great deal about 
their surface characteristics and their potential inter-actions 
3.3 INTERATOMIC BONDING 
Primary Bonds 
Only the outer shell or valence electrons participate in 
the formation of primary interatomic bonds. There are 
Copyrighted Material 
Figure 3.3 Simplified representation of an atom. 
with adjacent phases. 
Interatomic and intermolecular bonding forces hold 
matter together. Unbalanced forces exist at phase 
boundaries. The nature and magnitude of these forces 
influence the formation of soil minerals, the structure, 
size, and shape of soil particles, and the physicochem-ical 
phenomena that determine engineering properties 
and behavior. In this chapter some aspects of atomic 
and intermolecular forces, crystal structure, structure 
stability, and characteristics of surfaces that are perti-nent 
to the understanding of soil behavior are sum-marized 
simply and briefly. This is followed by a 
somewhat more detailed treatment of soil minerals and 
their characteristics. 
3.2 ATOMIC STRUCTURE 
Current concepts of atomic structure and interparticle 
bonding forces are based on quantum mechanics. An 
electron can have only certain values of energy. Elec-tronic 
energy can jump to a higher level by the ab-sorption 
of radiant energy or drop to a lower level by 
the emission of radiant energy. No more than two elec-trons 
in an atom can have the same energy level, and 
the spins of these two electrons must be in opposite 
directions. Different bonding characteristics for differ-ent 
elements exist because of the combined effects of 
electronic energy quantization and the limitation on the 
number of electrons at each energy level. 
An atom may be represented in simplified form by 
a small nucleus surrounded by diffuse concentric 
‘‘clouds’’ of electrons (Fig. 3.3). The maximum num-ber 
of electrons that may be located in each diffuse 
shell is determined by quantum theory. The number 
and arrangement of electrons in the outermost shell are 
of prime importance for the development of different 
types of interatomic bonding and crystal structure. 
Interatomic bonds form when electrons in adjacent 
atoms interact in such a way that their energy levels 
are lowered. If the energy reduction is large, then a 
strong, primary bond develops. The way in which the 
bonding electrons are localized in space determines 
whether or not the bonds are directional. The strength 
and directionality of interatomic bonds, together with 
the relative sizes of the bonded atoms, determine the 
type of crystal structure assumed by a given compo-sition. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
SECONDARY BONDS 39 
three limiting types: covalent, ionic, and metallic. They 
differ because of how the bonding electrons are local-ized 
in space. The energy of these bonds per mole of 
bonded atoms is from 60  103 to more than 400  
103 joules (J; 15 to 100 kcal). As there are 6.023  
1023 molecules per mole, it might be argued that such 
bonds are weak; however, relative to the weight of an 
atom they are very large. 
Covalent Bonds In the covalent bond, one or more 
bonding electrons are shared by two atomic nuclei to 
complete the outer shell for each atom. Covalent bonds 
are common in gases. If outer shell electrons are rep-resented 
Copyrighted Material 
by dots, then examples for (1) hydrogen gas, 
(2) methane, and (3) chlorine gas are: 
1. H  H  H:H 
H 
2. 
  
C  4H  H:C:H  H 
3. 
    
:Cl  Cl:  :Cl:Cl: 
In the solid state, covalent bonds form primarily be-tween 
nonmetallic atoms such as oxygen, chlorine, 
nitrogen, and fluorine. Since only certain electrons 
participate in the bonding, covalent bonds are direc-tional. 
As a result, atoms bonded covalently pack in 
such a way that there are fixed bond angles. 
Ionic Bonds Ionic bonds form between positively 
and negatively charged free ions that acquire their 
charge through gain or loss of electrons. Cations (pos-itively 
charged atoms that are attracted by the cathode 
in an electric field) form by atoms giving up one or 
more loosely held electrons that lie outside a com-pleted 
electron shell and have a high energy level. Met-als, 
alkalies (e.g., sodium, potassium), and alkaline 
earths (e.g., calcium, magnesium) form cations. Anions 
(negatively charged atoms that are attracted to the an-ode) 
are those atoms requiring only a few electrons to 
complete their outer shell. Because the outer shells of 
ions are complete, structures cannot form by electron 
sharing as in the case of the covalent bond. Since ions 
are electrically charged, however, strong electrical at-tractions 
(and repulsions) can develop between them. 
The ionic bond is nondirectional. Each cation at-tracts 
all neighboring anions. In sodium chloride, 
which is one of the best examples of ionic bonding, a 
sodium cation attracts as many chlorine anions as will 
fit around it. Geometric considerations and electrical 
neutrality determine the actual arrangement of ioni-cally 
bonded atoms. 
As ionic bonding causes a separation between the 
centers of positive and negative charge in a molecule, 
the molecule will orient in an electrical field forming 
a dipole. The strength of this dipole is expressed in 
terms of the dipole moment . If two electrical charges 
of magnitude e, where e is the electronic charge, 
are separated by a distance d, then 
  d  e (3.1) 
Covalently bonded atoms may also produce dipolar 
molecules. 
Metallic Bonds Metals contain loosely held val-ence 
electrons that hold the positive metal ions to-gether 
but are free to travel through the solid material. 
Metallic bonds are nondirectional and can exist only 
among a large group of atoms. It is the large group of 
electrons and their freedom to move that make metals 
such good conductors of electricity and heat. The me-tallic 
bond is of little importance in most soils. 
Bonding in Soil Minerals 
A combination of ionic and covalent bonding is typical 
in most nonmetallic solids. Purely ionic or covalent 
bonding is a limiting condition that is the exception 
rather than the rule in most cases. Silicate minerals are 
the most abundant constituents of most soils. The in-teratomic 
bond in silica (SiO2) is about half covalent 
and half ionic. 
3.4 SECONDARY BONDS 
Secondary bonds that are weak relative to ionic and 
covalent bonds also form between units of matter. They 
may be strong enough to determine the final arrange-ments 
of atoms in solids, and they may be sources of 
attraction between very small particles and between 
liquids and solid particles. 
The Hydrogen Bond 
If a hydrogen ion forms the positive end of a dipole, 
then its attraction to the negative end of an adjacent 
molecule is termed a hydrogen bond. Hydrogen bonds 
form only between strongly electronegative atoms such 
as oxygen and fluorine because these atoms produce 
the strongest dipoles. When the electron is detached 
from a hydrogen atom, such as when it combines with 
oxygen to form water, only a proton remains. As the 
electrons shared between the oxygen and hydrogen at-oms 
spend most of their time between the atoms, the 
oxygens act as the negative ends of dipoles, and the 
hydrogen protons act as the positive ends. The positive 
and negative ends of adjacent water molecules tie them 
together forming water and ice. 
The strength of the hydrogen bond is much greater 
than that of other secondary bonds because of the small 
size of the hydrogen ion. Hydrogen bonds are impor- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
40 3 SOIL MINERALOGY 
within the lattice where atoms or atomic groups 
Copyrighted Material 
Figure 3.4 Examples of some common crystals. (hkl) are 
cleavage plane indices. From Dana’s Manual of Mineralogy, 
by C. S. Hurlbut, 16th Edition. Copyright  1957 by John 
Wiley  Sons. Reprinted with permission from John Wiley 
 Sons. 
tant in determining some of the characteristics of the 
clay minerals and in the interaction between soil par-ticle 
surfaces and water. 
van der Waals Bonds 
Permanent dipole bonds such as the hydrogen bond are 
directional. Fluctuating dipole bonds, commonly 
termed van der Waals bonds, also exist because at any 
one time there may be more electrons on one side of 
the atomic nucleus than on the other. This creates weak 
instantaneous dipoles whose oppositely charged ends 
attract each other. 
Although individual van der Waals bonds are weak, 
typically an order of magnitude weaker than a hydro-gen 
bond, they are nondirectional and additive between 
atoms. Consequently, they decrease less rapidly with 
distance than primary valence and hydrogen bonds 
when there are large groups of atoms. They are strong 
enough to determine the final arrangements of groups 
of atoms in some solids (e.g., many polymers), and 
they may be responsible for small cohesions in fine-grained 
soils. Van der Waals forces are described fur-ther 
in Chapter 7. 
3.5 CRYSTALS AND THEIR PROPERTIES 
Particles composed of mineral crystals form the 
greatest proportion of the solid phase of a soil. A crys-tal 
is a homogeneous body bounded by smooth plane 
surfaces that are the external expression of an orderly 
internal atomic arrangement. A solid without internal 
atomic order is termed amorphous. 
Crystal Formation 
Crystals may form in three ways: 
1. From Solution Ions combine as they separate 
from solution and gradually build up a solid of 
definite structure and shape. Halite (sodium chlo-ride) 
and other evaporites are examples. 
2. By Fusion Crystals form directly from a liquid 
as a result of cooling. Examples are igneous rock 
minerals solidified from molten rock magma and 
ice from water. 
3. From Vapor Although not of particular impor-tance 
in the formation of soil minerals, crystals 
can form directly from cooling vapors. Examples 
include snowflakes and flowers of sulfur. 
Examples of some common crystals are shown in 
Fig. 3.4. 
Characteristics of Crystals 
Certain crystal characteristics are used to distinguish 
different classes or groups of minerals. Variations in 
these characteristics result in different properties. 
1. Structure The atoms in a crystal are arranged 
in a definite orderly manner to form a three-dimensional 
network termed a lattice. Positions 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
CRYSTALS AND THEIR PROPERTIES 41 
are located are termed lattice points. Only 14 dif-ferent 
arrangements of lattice points in space are 
possible. These are the Bravais space lattices, 
and they are illustrated in Fig. 3.5. 
The smallest subdivision of a crystal that still 
possesses the characteristic composition and spa-tial 
same substance are constant. Crystals break 
along smooth cleavage planes. Cleavage planes 
lie between planes in which the atoms are most 
Copyrighted Material 
arrangement of atoms in the crystal is the unit 
cell. The unit cell is the basic repeating unit of 
the space lattice. 
2. Cleavage and Outward Form The angles be-tween 
corresponding faces on crystals of the 
Figure 3.5 Unit cells of the 14 Bravais space lattices. The capital letters refer to the type 
of cell: P, primitive cell; C, cell with a lattice point in the center of two parallel faces; F, 
cell with a lattice point in the center of each face; I, cell with a lattice point in the center 
of the interior; R, rhombohedral primitive cell. All points indicated are lattice points. There 
is no general agreement on the unit cell to use for the hexagonal Bravais lattice; some prefer 
the P cell shown with solid lines, and others prefer the C cell shown in dashed lines (modified 
from Moffatt et al., 1965). 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
42 3 SOIL MINERALOGY 
Copyrighted Material 
Figure 3.6 The six crystal systems. 
densely packed. This is because the center-to-center 
distance between atoms on opposite sides 
of the plane is greater than along other planes 
through the crystal. As a result, the strength along 
cleavage planes is less than in other directions. 
3. Optical Properties The specific atomic arrange-ments 
within crystals allow light diffraction and 
polarization. These properties are useful for iden-tification 
and classification. Identification of rock 
minerals by optical means is common. Optical 
studies in soil are less useful because of the small 
sizes of most soil particles. 
4. X-ray and Electron Diffraction The orderly 
atomic arrangements in crystals cause them to 
behave with respect to X-ray and electron beams 
in much the same way as does a diffraction grat-ing 
with respect to visible light. Different crystals 
yield different diffraction patterns. This makes X-ray 
diffraction a powerful tool for the study and 
identification of very small particles, such as clay 
that cannot be seen using optical means. 
5. Symmetry There are 32 distinct crystal classes 
based on symmetry considerations involving the 
arrangement and orientation of crystal faces. 
These 32 classes may be grouped into 6 crystal 
systems with the classes within each system bear-ing 
close relationships to each other. 
The six crystal systems are illustrated in Fig. 3.6. 
Crystallographic axes parallel to the intersection edges 
of prominent crystal faces are established for each of 
the six crystal systems. In most crystals, these axes will 
also be symmetry axes or axes normal to symmetry 
planes. In five of the six systems, the crystals are re-ferred 
to three crystallographic axes. In the sixth (the 
hexagonal system), four axes are used. The axes are 
denoted by a, b, c (a1, a2, a3, and c in the hexagonal 
system) and the angles between the axes by , , 
and 	. 
Isometric or Cubic System There are three mutu-ally 
perpendicular axes of equal length. Mineral 
examples are galena, halite, magnetite, and pyrite. 
Hexagonal System Three equal horizontal axes ly-ing 
in the same plane intersect at 60 with a fourth 
axis perpendicular to the other three and of dif-ferent 
length. Examples are quartz, brucite, cal-cite, 
and beryl. 
Tetragonal System There are three mutually per-pendicular 
axes, with two horizontal of equal 
length, but different than that of the vertical axis. 
Zircon is an example. 
Orthorhombic System There are three mutually 
perpendicular axes, each of different length. Ex-amples 
include sulfur, anhydrite, barite, diaspore, 
and topaz. 
Monoclinic System There are three unequal axes, 
two inclined to each other at an oblique angle, 
with the third perpendicular to the other two. Ex-amples 
are orthoclase feldspar, gypsum, musco-vite, 
biotite, gibbsite, and chlorite. 
Triclinic System Three unequal axes intersect at 
oblique angles. Examples are plagioclase feldspar, 
kaolinite, albite, microcline, and turquoise. 
3.6 CRYSTAL NOTATION 
Miller indices are used to describe plane orientations 
and directions in a crystal. This information, along 
with the distances that separate parallel planes is im-portant 
for the identification and classification of dif-ferent 
minerals. All lengths are expressed in terms of 
unit cell lengths. Any plane through a crystal may be 
described by intercepts, in terms of unit cell lengths, 
on the three or four crystallographic axes for the sys-tem 
in which the crystal falls. The reciprocals of these 
intercepts are used to index the plane. Reciprocals are 
used to avoid fractions and to account for planes par-allel 
to an axis (an intercept of infinity equals an index 
value of 0). 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
CRYSTAL NOTATION 43 
lengths. Take plane mnp in Fig. 3.7a as an example. 
The intercepts of this plane are a  1, b  1, and 
c  1. The Miller indices of this plane are found by 
taking the reciprocals of these intercepts and clearing 
of fractions. Thus, 
Reciprocals are 1/1, 1/1, 1/1 
Miller indices are (111) 
The indices are always enclosed within parentheses 
and indicated in the order abc without commas. Paren- 
Copyrighted Material 
An example illustrates the determination and mean-ing 
of Miller indices. Consider the mineral muscovite, 
a member of the monoclinic system. It has unit cell 
dimensions of a  0.52 nanometers (nm), b  0.90 
nm, c  2.0 nm, and   95 30. Both the compo-sition 
and crystal structure of muscovite are similar to 
those of some of the important clay minerals. 
The muscovite unit cell dimensions and intercepts 
are shown in Fig. 3.7a. The intercepts for any plane of 
interest are first determined in terms of unit cell 
Figure 3.7 Miller indices: (a) Unit cell of muscovite, (b) (002) plane for muscovite, (c) 
(014) plane for muscovite, and (d) (623) plane for muscovite. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
44 3 SOIL MINERALOGY 
Table 3.1 Atomic Packing, Structure, and Structural 
Stability 
Radius 
Ratioa Nb Geometry Example Stability 
0–0.155 2 Line — — 
0.155– 
0.225 
3 Triangle (CO3)2 Very high 
Material 
0.225– 
0.414 
Copyrighted 4 Tetrahedron (SiO4)4 Moderately 
high 
0.414– 
0.732 
6 Octahedron [Al(OH)6]3 High 
0.732– 
1.0 
8 Body-cen-tered 
cube 
Iron Low 
1.0 12 Sheet K–O bond 
in mica 
Very low 
aRange of cation to anion diameter ratios over which 
stable coordination is expected. 
bCoordination number. 
Table 3.2 Relative Stabilities of Some Soil Mineral 
Structural Units 
Structural Unit 
Approximate 
Relative Bond 
Strength 
(Valence/N) 
Silicon tetrahedron, (SiO4)4 4/4  1 
Aluminum tetrahedron, [Al(OH)4]1 3/4 
Aluminum octahedron, [Al(OH)6]3 3/6  1/2 
Magnesium octahedron, [Mg(OH)6]4 2/6  1/3 
K–O12 
23 1/12 
theses are always used to indicate crystallographic 
planes, whereas brackets are used to indicate direc-tions. 
For example, [111] designates line oq in Fig. 
3.7a. Additional examples of Miller indices for planes 
through the muscovite crystal are shown in Figs. 3.7b, 
3.7c, and 3.7d. A plane that cuts a negative axis is 
designated by placing a bar over the index that pertains 
to the negative intercept (Fig. 3.7d). The general index 
(hkl) is used to refer to any plane that cuts all three 
axes. Similarly (h00) designates a plane cutting only 
the a axis, (h0l) designates a plane parallel to the b 
axis, and so on. For crystals in the hexagonal system, 
the Miller index contains four numbers. The (001) 
planes of soil minerals are of particular interest be-cause 
they are indicative of specific clay mineral types. 
3.7 FACTORS CONTROLLING CRYSTAL 
STRUCTURES 
Organized crystal structures do not develop by chance. 
The most stable arrangement of atoms in a crystal is 
that which minimizes the energy per unit volume. This 
is achieved by preserving electrical neutrality, satisfy-ing 
bond directionality, minimizing strong ion repul-sions, 
and packing atoms closely together. 
If the interatomic bonding is nondirectional, then the 
relative atomic sizes have a controlling influence on 
packing. The closest possible packing will maximize 
the number of bonds per unit volume and minimize the 
bonding energy. If interatomic bonds are directional, 
as is the case for covalent bonds, then both bond angles 
and atomic size are important. 
Anions are usually larger than cations because of 
electron transfer from cations to anions. The number 
of nearest neighbor anions that a cation possesses in a 
structure is termed the coordination number (N) or li-gancy. 
Possible values of coordination number in solid 
structures are 1 (trivial), 2, 3, 4, 6, 8, and 12. The 
relationships between atomic sizes, expressed as the 
ratio of cationic to anionic radii, coordination number, 
and the geometry formed by the anions are indicated 
in Table 3.1. 
Most solids do not have bonds that are completely 
nondirectional, and the second nearest neighbors may 
influence packing as well as the nearest neighbors. 
Even so, the predicted and observed coordination num-bers 
are in quite good agreement for many materials. 
The valence of the cation divided by the number of 
coordinated anions is an approximate indication of the 
relative bond strength, which, in turn, is related to the 
structural stability of the unit. Some of the structural 
units common in soil minerals and their relative bond 
strengths are listed in Table 3.2. 
The basic coordination polyhedra are seldom elec-trically 
neutral. In crystals formed by ionic bonded pol-yhedra, 
the packing maintains electrical neutrality and 
minimizes strong repulsions between ions with like 
charge. In such cases, the valence of the central cation 
equals the total charge of the coordinated anions, and 
the unit is really a molecule. Units of this type are held 
together by weaker, secondary bonds. An example is 
brucite, a mineral that has the composition Mg(OH)2. 
The Mg2 ions are in octahedral coordination with six 
(OH) ions forming a sheet structure in such a way 
that each (OH) is shared by 3Mg2. In a sheet con-taining 
N Mg2 ions, therefore, there must be 6N/3  
2N (OH) ions. Thus, electrical neutrality results, and 
the sheet is in reality a large molecule. Successive oc- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
SURFACES 45 
tahedral sheets are loosely bonded by van der Waals 
forces. Because of this, brucite has perfect basal cleav-age 
parallel to the sheets. 
Cations concentrate their charge in a smaller volume 
than do anions, so the repulsion between cations is 
greater than between anions. Cationic repulsions are 
minimized when the anions are located at the centers 
of coordination polyhedra. If the cations have a low 
valence, then the anion polyhedra pack as closely as 
possible to minimize energy per unit volume. If, on the 
other hand, the cations are small and highly charged, 
then the units arrange in a variety of ways in response 
to the repulsions. The silicon cation is in this category. 
Copyrighted Material 
3.8 SILICATE CRYSTALS 
Small cations form structures with coordination num-bers 
of 3 and 4 (Table 3.1). These cations are often 
highly charged and generate strong repulsions between 
adjacent triangles or tetrahedra. As a result, such struc-tures 
share only corners and possibly edges, but never 
faces, since to do so would bring the cations too close 
together. The radius of silicon is only 0.039 nm, 
whereas that of oxygen is 0.132 nm. Thus silicon and 
oxygen combine in tetrahedral coordination, with the 
silicon occupying the space at the center of the tetra-hedron 
formed by the four oxygens. The tetrahedral 
arrangement satisfies both the directionality of the 
bonds (the Si–O bond is about half covalent and half 
ionic) and the geometry imposed by the radius ratio. 
Silicon is very abundant in Earth’s crust, amounting to 
about 25 percent by weight, but only 0.8 percent by 
volume. Almost half of igneous rock by weight and 
91.8 percent by volume is oxygen. 
Silica tetrahedra join only at their corners, and 
sometimes not at all. Thus many crystal structures are 
possible, and there is a large number of silicate min-erals. 
Silicate minerals are classified according to how 
the silica tetrahedra (SiO4)4 associate with each other, 
as shown in Fig. 3.8. The tetrahedral combinations in-crease 
in complexity from the beginning to the end of 
the figure. The structural stability increases in the same 
direction. 
Island (independent) silicates are those in which the 
tetrahedra are not joined to each other. Instead, the four 
excess oxygen electrons are bonded to other positive 
ions in the crystal structure. In the olivine group, the 
minerals have the composition R2 2 
 SiO4. Garnets 
4 
contain cations of different valences and coordination 
numbers R2  R3(SiO). The negative charge of the 
3 
2 
43SiO4 group in zircon is all balanced by the single Zr4. 
Ring and chain silicates are formed when corners of 
tetrahedra are shared. The formulas for these structures 
contain (SiO3)2. The pyroxene minerals are in this 
class. Enstatite, MgSiO3, is a simple member of this 
group. Some of the positions normally occupied by 
Si4 in single-chain structures may be filled by Al3. 
Substitution of ions of one kind by ions of another 
type, having either the same or different valence, but 
the same crystal structure, is termed isomorphous sub-stitution. 
The term substitution implies a replacement 
whereby a cation in the structure is replaced at some 
time by a cation of another type. In reality, however, 
the replaced cations were never there, and the mineral 
was formed with its present proportions of the different 
cations in the structure. 
Double chains of indefinite length may form with 
(Si4O11)6 as part of the structure. The amphiboles fall 
into this group (Fig. 3.8). Hornblendes have the same 
basic structure, but some of the Si4 positions are filled 
by Al3. The cations Na and K can be incorporated 
into the structure to satisfy electrical neutrality; Al3, 
Fe3, Fe2, and Mn2 can replace part of the Mg2 in 
sixfold coordination, and the (OH) group can be re-placed 
by F. 
In sheet silicates three of the four oxygens of each 
tetrahedron are shared to give structures containing 
(Si2O5)2. The micas, chlorites, and many of the clay 
minerals contain silica in a sheet structure. Framework 
silicates result when all four of the oxygens are shared 
with other tetrahedra. The most common example is 
quartz. In quartz, the silica tetrahedra are grouped to 
form spirals. The feldspars also have three-dimensional 
framework structures. Some of the silicon positions are 
filled by aluminum, and the excess negative charge 
thus created is balanced by cations of high coordina-tion 
such as potassium, calcium, sodium, and barium. 
Differences in the amounts of this isomorphous sub-stitution 
are responsible for the different members of 
the feldspar family. 
3.9 SURFACES 
All liquids and solids terminate at a surface, or phase 
boundary, on the other side of which is matter of a 
different composition or state. In solids, atoms are 
bonded into a three-dimensional structure, and the ter-mination 
of this structure at a surface, or phase bound-ary, 
produces unsatisfied force fields. In a fine-grained 
particulate material such as clay soil the surface area 
may be very large relative to the mass of the material, 
and, as is emphasized throughout this book, the influ-ences 
of the surface forces on properties and behavior 
may be very large. 
Unsatisfied forces at solid surfaces may be balanced 
in any of the following ways: 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
46 3 SOIL MINERALOGY 
Copyrighted Material 
Figure 3.8 Silica tetrahedral arrangements in different silicate mineral structures. Reprinted 
Gillott (1968) with permission from Elsevier Science Publishers BV. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
SURFACES 47 
Material 
Copyrighted Figure 3.8 (Continued) 
1. Attraction and adsorption of molecules from the 
adjacent phase 
2. Cohesion with the surface of another mass of the 
same substance 
3. Solid-state adjustments of the structure beneath 
the surface. 
Each unsatisfied bond force is significant relative to 
the weight of atoms and molecules. The actual mag-nitude 
of 1011 N or less, however, is infinitesimal 
compared to the weight of a piece of gravel or a grain 
of sand. On the other hand, consider the effect of re-ducing 
particle size. A cube 10 mm on an edge has a 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
48 3 SOIL MINERALOGY 
surface area of 6.0  104 m2. If it is cut in half in 
the three directions, eight cubes result, each 5 mm on 
an edge. The surface area now is 12.0  104 m2. If 
the cubes are further divided to 1 m on an edge, the 
surface becomes 6.0 m2 for the same 1000 mm3 of 
material. Thus, as a solid is subdivided into smaller 
and smaller units, the proportion of surface area to 
weight becomes larger and larger. For a given particle 
shape, the ratio of surface area to volume is inversely 
proportional to some effective particle diameter. 
For many materials when particle size is reduced to 
1 or 2 m or less the surface forces begin to exert a 
distinct influence on the behavior. Study of the behav-ior 
Copyrighted Material 
of particles of this size and less requires consider-ations 
of colloidal and surface chemistry. Most clay 
particles behave as colloids, both because of their 
small size and because they have unbalanced surface 
electrical forces as a result of isomorphous substitu-tions 
within their structure. 
Montmorillonite, which is one of the members of 
the smectite clay mineral group (see Section 3.17), 
may break down into particles that are only 1 unit cell 
thick (1.0 nm) when in a dispersed state and have a 
specific surface area of 800 m2 /g. If all particles con-tained 
in about 10 g of this clay could be spread out 
side by side, they would cover a football field. 
3.10 GRAVEL, SAND, AND SILT PARTICLES 
The physical characteristics of cohesionless soils, that 
is, gravel, sand, and nonplastic silts, are determined 
primarily by particle size, shape, surface texture, and 
size distribution. The mineral composition determines 
hardness, cleavage, and resistance to physical and 
chemical breakdown. Some carbonate and sulfate min-erals, 
such as calcite and gypsum, are sufficiently sol-uble 
that their decomposition may be significant within 
the time frame of many projects. In many cases, how-ever, 
the nonclay particles may be treated as relatively 
inert, with interactions that are predominantly physical 
in nature. Evidence of this is provided by the soils on 
the Moon. Lunar soils have a silty, fine sand gradation; 
however, their compositions are totally different than 
those of terrestrial soils of the same gradation. The 
engineering properties of the two materials are sur-prisingly 
similar, however. 
The gravel, sand, and most of the silt fraction in a 
soil are composed of bulky, nonclay particles. As most 
soils are the products of the breakdown of preexisting 
rocks and soils, they are weathering products. Thus, 
the predominant mineral constituents of any soil are 
those that are one or more of the following: 
1. Very abundant in the source material 
2. Highly resistant to weathering, abrasion, and im-pact 
3. Weathering products 
The nonclays are predominantly rock fragments or 
mineral grains of the common rock-forming minerals. 
In igneous rocks, which are the original source mate-rial 
for many soils, the most prevalent minerals are the 
feldspars (about 60 percent) and the pyroxenes and 
amphiboles (about 17 percent). Quartz accounts for 
about 12 percent of these rocks, micas for 4 percent, 
and other minerals for about 8 percent. 
However, in most soils, quartz is by far the most 
abundant mineral, with small amounts of feldspar and 
mica also present. Pyroxenes and amphiboles are sel-dom 
found in significant amounts. Carbonate minerals, 
mainly calcite and dolomite, are also found in some 
soils and can occur as bulky particles, shells, precipi-tates, 
or in solution. Carbonates dominate the compo-sition 
of some deep-sea sediments. Sulfates, in various 
forms, are found primarily in soils of semiarid and arid 
regions, with gypsum (CaSO4  2H2O) being the most 
common. Iron and aluminum oxides are abundant in 
residual soils of tropical regions. 
Quartz is composed of silica tetrahedra grouped to 
form spirals, with all tetrahedral oxygens bonded to 
silicon. The tetrahedral structure has a high stability. 
In addition, the spiral grouping of tetrahedra produces 
a structure without cleavage planes, quartz is already 
an oxide, there are no weakly bonded ions in the struc-ture, 
and the mineral has high hardness. Collectively, 
these factors account for the high persistence of quartz 
in soils. 
Feldspars are silicate minerals with a three-dimensional 
framework structure in which part of the 
silicon is replaced by aluminum. The excess negative 
charge resulting from this replacement is balanced by 
cations such as potassium, calcium, sodium, strontium, 
and barium. As these cations are relatively large, their 
coordination number is also large. This results in an 
open structure with low bond strengths between units. 
Consequently, there are cleavage planes, the hardness 
is only moderate, and feldspars are relatively easily 
broken down. This accounts for their lack of abun-dance 
in soils compared to their abundance in igneous 
rocks. 
Mica has a sheet structure composed of tetrahedral 
and octahedral units. Sheets are stacked one on the 
other and held together primarily by potassium ions in 
12-fold coordination that provide an electrostatic bond 
of moderate strength. In comparison with the intralayer 
bonds, however, this bond is weak, which accounts for 
the perfect basal cleavage of mica. As a result of the 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
STRUCTURAL UNITS OF THE LAYER SILICATES 49 
thin-plate morphology of mica flakes, sand and silts 
containing only a few percent mica may exhibit high 
compressibility when loaded and large swelling when 
unloaded, as may be seen in Fig. 3.9. The amphiboles, 
pyroxenes, and olivine have crystal structures that are 
rapidly broken down by weathering; hence they are 
absent from most soils. 
Some examples of silt and sand particles from dif-ferent 
Copyrighted Material 
soils are shown in Fig. 3.10. Angularity and 
roundness can be used to describe particle shapes, as 
shown in Fig. 3.11. Elongated and platy particles can 
develop preferred orientations, which can be respon-sible 
for anisotropic properties within a soil mass. The 
surface texture of the grains influences the stress– 
deformation and strength properties. 
3.11 SOIL MINERALS AND MATERIALS 
FORMED BY BIOGENIC AND GEOCHEMICAL 
PROCESSES 
Evaporite deposits formed by precipitation of salts 
from salt lakes and seas as a result of the evaporation 
of water are sometimes found in layers that are several 
meters thick. The major constituents of seawater and 
their relative proportions are listed in Table 3.3. Also 
listed are some of the more important evaporite de-posits. 
Figure 3.9 Swelling index as a function of mica content for 
coarse-grained mixtures (data from Terzaghi, 1931). 
In some areas alternating layers of evaporite and 
clay or other fine-grained sediments are formed during 
cyclic wet and dry periods. 
Many limestones, as well as coral, have been formed 
by precipitation or from the remains of various organ-isms. 
Because of the much greater solubility of lime-stone 
than most other rock types, it may be the source 
of special problems caused by solution channels and 
cavities under foundations. 
Chemical sediments and rocks in freshwater lakes, 
ponds, swamps, and bays are occasionally encountered 
in civil engineering projects. Biochemical processes 
form marl, which ranges from relatively pure calcium 
carbonate to mixtures with mud and organic matter. 
Iron oxide is formed in some lakes. Diatomite or dia-tomaceous 
earth is essentially pure silica formed from 
the skeletal remains of small (up to a few tenths of a 
millimeter) freshwater and saltwater organisms. Owing 
to their solubility limestone, calcite, gypsum, and other 
salts may cause special geotechnical problems. 
Oxidation and reduction of pyrite-bearing earth ma-terials, 
that is, soils and rocks containing FeS2, can be 
the source of many types of geotechnical problems, 
including ground heave, high swell pressures, forma-tion 
of acid drainage, damage to concrete, and corro-sion 
of steel (Bryant et al., 2003). The chemical and 
biological processes and consequences of pyritic re-actions 
are covered in Sections 8.3, 8.11, and 8.16. 
More than 12 percent of Canada is covered by a 
peaty material, termed muskeg, composed almost en-tirely 
of decaying vegetation. Peat and muskeg may 
have water contents of 1000 percent or more; they are 
very compressible, and they have low strength. The 
special properties of these materials and methods for 
analysis of geotechnical problems associated with 
them are given by MacFarlane (1969), Dhowian and 
Edil (1980), and Edil and Mochtar (1984). 
3.12 SUMMARY OF NONCLAY MINERAL 
CHARACTERISTICS 
Important compositional, structural, and morphological 
characteristics of the important nonclay minerals found 
in soils are summarized in Table 3.4. Of these miner-als, 
quartz is by far the most common, both in terms 
of the number of soils in which it is found and its 
abundance in a typical soil. Feldspar and mica are fre-quently 
present in small percentages. 
3.13 STRUCTURAL UNITS OF THE LAYER 
SILICATES 
Clay minerals in soils belong to the mineral family 
termed phyllosilicates, which also contains other layer 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
50 3 SOIL MINERALOGY 
Material 
Copyrighted Figure 3.10 Photomicrographs of sand and silt particles from several soils: (a) Ottawa stan-dard 
sand, (b) Monterey sand, (c) Sacramento River sand, (d) Eliot sand, and (e) lunar soil 
mineral grains (photo courtesy Johnson Space Center). Squares in background area are 11 
mm. (ƒ) Recrystallized breccia particles from lunar soil (photo courtesy of NASA Johnson 
Space Center). Squares in background grid are 11 mm. 
silicates such as serpentine, pyrophyllite, talc, mica, 
and chlorite. Clay minerals occur in small particle 
sizes, and their unit cells ordinarily have a residual 
negative charge that is balanced by the adsorption of 
cations from solution. 
The structures of the common layer silicates are 
made up of combinations of two simple structural 
units, the silicon tetrahedron (Fig. 3.12) and the alu-minum 
or magnesium octahedron (Fig. 3.13). Different 
clay mineral groups are characterized by the stacking 
arrangements of sheets1 (sometimes chains) of these 
1 In conformity with the nomenclature of the Clay Minerals Society 
(Bailey et al., 1971), the following terms are used: a plane of atoms, 
a sheet of basic structural units, and a layer of unit cells composed 
of two, three, or four sheets. 
units and the manner in which two successive two- or 
three-sheet layers are held together. 
Differences among minerals within clay mineral 
groups result primarily from differences in the type and 
amount of isomorphous substitution within the crystal 
structure. Possible substitutions are nearly endless in 
number, and the crystal structure arrangement may 
range from very poor to nearly perfect. Fortunately for 
engineering purposes, knowledge of the structural and 
compositional characteristics of each group, without 
detailed study of the subtleties of each specific mineral, 
is adequate. 
Silica Sheet 
In most clay mineral structures, the silica tetrahedra 
are interconnected in a sheet structure. Three of the 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
STRUCTURAL UNITS OF THE LAYER SILICATES 51 
Material 
Copyrighted Figure 3.10 (Continued) 
Figure 3.11 Sand and silt size particle shapes as seen in 
silhouette. 
four oxygens in each tetrahedron are shared to form a 
hexagonal net, as shown in Figs. 3.12b and 3.14. The 
bases of the tetrahedra are all in the same plane, and 
the tips all point in the same direction. The structure 
has the composition (Si4O10)4 and can repeat indefi-nitely. 
Electrical neutrality can be obtained by replace-ment 
of four oxygens by hydroxyls or by union with 
a sheet of different composition that is positively 
charged. The oxygen-to-oxygen distance is 2.55 ang-stroms 
(A˚ ),2 the space available for the silicon ion is 
0.55 A˚ 
, and the thickness of the sheet in clay mineral 
structures is 4.63 A˚ 
(Grim, 1968). 
2 In conformity with the SI system of units, lengths should be given 
in nanometers. For convenience, however, the angstrom unit is re-tained 
for atomic dimensions, where 1 A˚ 
 0.1 nm. 
Silica Chains 
In some of the less common clay minerals, silica tet-rahedra 
are arranged in bands made of double chains 
of composition (Si4O11)6. Electrical neutrality is 
achieved and the bands are bound together by alumi-num 
and/or magnesium ions. A diagrammatic sketch 
of this structure is shown in Fig. 3.8. Minerals in this 
group resemble the amphiboles in structure. 
Octahedral Sheet 
This sheet structure is composed of magnesium or alu-minum 
in octahedral coordination with oxygens or hy-droxyls. 
In some cases, other cations are present in 
place of Al3 and Mg2, such as Fe2, Fe3, Mn2, 
Ti4, Ni2, Cr3, and Li. Figure 3.13b is a schematic 
diagram of such a sheet structure. The oxygen-to-oxygen 
distance is 2.60 A˚ 
, and the space available for 
the octahedrally coordinated cation is 0.61 A˚ 
. The 
thickness of the sheet is 5.05 A˚ 
in clays (Grim, 1968). 
If the cation is trivalent, then normally only two-thirds 
of the possible cationic spaces are filled, and the 
structure is termed dioctahedral. In the case of alu-minum, 
the composition is Al2(OH)6. This composition 
and structure form the mineral gibbsite. When com-bined 
with silica sheets, as is the case in clay mineral 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
52 3 SOIL MINERALOGY 
Table 3.3 Major Constituents of Seawater and Evaporite Deposits 
Ion Grams per Liter 
Percent by Weight 
of Total Solids Important Evaporite Deposits 
Sodium, Na 10.56 30.61 Anhydrite CaSO4 
Magnesium, Mg2 1.27 3.69 Barite BaSO4 
Calcium, Ca2 0.40 1.16 Celesite SrSO4 
Potassium, K 0.38 1.10 Kieserite MgSO4  H2O 
Strontium, Sr2 0.013 0.04 Gypsum CaSO4  2H2O 
Chloride, Cl 18.98 55.04 Polyhalite Ca2K2Mg(SO4)  2H2O 
Sulfate, SO4 
Copyrighted Material 
2 2.65 7.68 Bloedite Ma2Mg(SO4)2  4H2O 
Bicarbonate, HCO3 
 0.14 0.41 Hexahydrite MgSO4  6H2O 
Bromide, Br 0.065 0.19 Epsomite MgSO4  7H2O 
Fluoride, F 0.001 — Kainite K4Mg4(Cl/SO4)  1 1H2 
O 
Boric Acid, H3BO3 0.026 0.08 Halite NaCl 
34.485 100.00 Sylvite KCl 
Flourite CaF2 
Bischofite MgCl2  6H2O 
Carnallite KMgCl3  6H2O 
Adapted from data by Degens (1965). 
structures, an aluminum octahedral sheet is referred to 
as a gibbsite sheet. 
If the octahedrally coordinated cation is divalent, 
then normally all possible cation sites are occupied and 
the structure is trioctahedral. In the case of magne-sium, 
the composition is Mg3(OH)6, giving the mineral 
brucite. In clay mineral structures, a sheet of magne-sium 
octahedra is termed a brucite sheet. 
Schematic representations of the sheets are useful 
for simplified diagrams of the structures of the differ-ent 
clay minerals: 
Silica sheet or 
Octahedral sheet (Various cations in octahedral coordination) 
Gibbsite sheet (Octahedral sheet cations are mainly aluminum) 
Brucite sheet (Octahedral sheet cations are mainly magnesium) 
Water layers are found in some structures and may 
be represented by  for each molecular layer. 
Atoms of a specific type, for example, potassium, are 
represented thus: K . 
The diagrams are indicative of the clay mineral layer 
structure. They do not indicate the correct width-to-length 
ratios for the actual particles. The structures 
shown are idealized; in actual minerals, irregular sub-stitutions 
and interlayering or mixed-layer structures 
are common. Furthermore, direct assembly of the basic 
units does not necessarily form the naturally occurring 
minerals. The ‘‘building block’’ approach is useful, 
however, for the development of conceptual models. 
3.14 SYNTHESIS PATTERN AND 
CLASSIFICATION OF THE CLAY MINERALS 
The manner in which atoms are assembled into tetra-hedral 
and octahedral units, followed by the formation 
of sheets and their stacking to form layers that combine 
to produce the different clay mineral groups is illus-trated 
in Fig. 3.15. The basic structures shown in the 
bottom row of Fig. 3.15 comprise the great prepon-derance 
of the clay mineral types that are found in 
soils. 
Grouping the clay minerals according to crystal 
structure and stacking sequence of the layers is con-venient 
since members of the same group have gen- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
SYNTHESIS PATTERN AND CLASSIFICATION OF THE CLAY MINERALS 53 
Table 3.4 Properties and Characteristics of Nonclay Minerals in Soils 
Mineral Formula 
Crystal 
System Cleavage 
Particle 
Shape 
Specific 
Gravity Hardness 
Occurrence 
in Soils of 
Engineering 
Interest 
Quartz SiO2 Hexagonal None Bulky 2.65 7 Very 
abundant 
Orthoclase 
feldspar 
Copyrighted Material 
KalSi3O8 Monoclinic 2 planes Elongate 2.57 6 Common 
Plagioclase 
feldspar 
NaAlSi3O8 
CaAl2Si3O8 (variable) 
Triclinic 2 planes Bulky— 
elongate 
2.62–2.76 6 Common 
Muscovite 
mica 
Kal3Si3O10(OH)2 Monoclinic Perfect basal Thin plates 2.76–3.1 2–21⁄2 Common 
Biotite mica K(Mg,FE)3AlSi3O10(OH)2 Monoclinic Perfect basal Thin plates 2.8–3.2 21⁄2–3 Common 
Hornblende Na,Ca,Mg,Fe,Al silicate Monoclinic Perfect 
prismatic 
Prismatic 3.2 5–6 Uncommon 
Augite 
(pyroxene) 
Ca(Mg,Fe,Al)(Al,Si)2O6 Monoclinic Good prismatic Prismatic 3.2–3.4 5–6 Uncommon 
Olivine (Mg,Fe)2SiO4 Orthorhombic Conchoidal 
fracture 
Bulky 3.27–3.37 61⁄2–7 Uncommon 
Calcite CaCO3 Hexagonal Perfect Bulky 2.72 21⁄2–3 May be 
abundant 
locally 
Dolomite CaMg(CO3)2 Hexagonal Perfect 
rhombohedral 
Bulky 2.85 31⁄2–4 May be 
abundant 
locally 
Gypsum CaSO4  2H2O Monoclinic 4 planes Elongate 2.32 2 May be 
abundant 
locally 
Pyrite FeS2 Isometric Cubical Bulky cubic 5.02 6–61⁄2 
Data from Hurlbut (1957). 
Figure 3.12 Silicon tetrahedron and silica tetrahedra arranged in a hexagonal network. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
54 3 SOIL MINERALOGY 
Figure 3.13 Octahedral unit and sheet Material 
structure of octahedral units. 
Copyrighted Figure 3.14 Silica sheet in plan view. 
erally similar engineering properties. The minerals 
have unit cells consisting of two, three, or four sheets. 
The two-sheet minerals are made up of a silica sheet 
and an octahedral sheet. The unit layer of the three-sheet 
minerals is composed of either a dioctahedral or 
trioctahedral sheet sandwiched between two silica 
sheets. Unit layers may be stacked closely together or 
water layers may intervene. The four-sheet structure of 
chlorite is composed of a 21 layer plus an interlayer 
hydroxide sheet. In some soils, inorganic, claylike ma-terial 
is found that has no clearly identifiable crystal 
structure. Such material is referred to as allophane or 
noncrystalline clay. 
The bottom row of Fig. 3.15 shows that the 21 
minerals differ from each other mainly in the type and 
amount of ‘‘glue’’ that holds the successive layers to-gether. 
For example, smectite has loosely held cations 
between the layers, illite contains firmly fixed potas-sium 
ions, and vermiculite has somewhat organized 
layers of water and cations. The chlorite group repre-sents 
an end member that has 21 layers bonded by an 
organized hydroxide sheet. The charge per formula 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
INTERSHEET AND INTERLAYER BONDING IN THE CLAY MINERALS 55 
Material 
Figure 3.15 Synthesis pattern for the clay minerals. 
Copyrighted unit is variable both within and among groups, and 
reflects the fact that the range of compositions is great 
owing to varying amounts of isomorphous substitution. 
Accordingly, the boundaries between groups are some-what 
arbitrary. 
Isomorphous Substitution 
The concept of isomorphous substitution was intro-duced 
in Section 3.13 in connection with some of the 
silicate crystals. It is very important in the structure 
and properties of the clay minerals. In an ideal gibbsite 
sheet, only two-thirds of the octahedral positions are 
filled, and all of the cations are aluminum. In an ideal 
brucite sheet, all the octahedral spaces are filled by 
magnesium. In an ideal silica sheet, silicons occupy all 
tetrahedral spaces. In clay minerals, however, some of 
the tetrahedral and octahedral spaces are occupied by 
cations other than those in the ideal structure. Common 
examples are aluminum in place of silicon, magnesium 
instead of aluminum, and ferrous iron (Fe2) for mag-nesium. 
This presence in an octahedral or tetrahedral 
position of a cation other than that normally found, 
without change in crystal structure, is isomorphous 
substitution. The actual tetrahedral and octahedral cat-ion 
distributions may develop during initial formation 
or subsequent alteration of the mineral. 
3.15 INTERSHEET AND INTERLAYER 
BONDING IN THE CLAY MINERALS 
A single plane of atoms that are common to both the 
tetrahedral and octahedral sheets forms a part of the 
clay mineral layers. Bonding between these sheets is 
of the primary valence type and is very strong. How-ever, 
the bonds holding the unit layers together may 
be of several types, and they may be sufficiently weak 
that the physical and chemical behavior of the clay is 
influenced by the response of these bonds to changes 
in environmental conditions. 
Isomorphous substitution in all of the clay minerals, 
with the possible exception of those in the kaolinite 
group, gives clay particles a net negative charge. To 
preserve electrical neutrality, cations are attracted and 
held between the layers and on the surfaces and edges 
of the particles. Many of these cations are exchange-able 
cations because they may be replaced by cations 
of another type. The quantity of exchangeable cations 
is termed the cation exchange capacity (cec) and is 
usually expressed as milliequivalents (meq)3 per 100 g 
of dry clay. 
Five types of interlayer bonding are possible in the 
layer silicates (Marshall, 1964). 
1. Neutral parallel layers are held by van der Waals 
forces. Bonding is weak; however, stable crystals 
of appreciable thickness such as the nonclay min- 
3Equivalent weight  combining weight of an element  (atomic 
weight /valence). Number of equivalents  (weight of element / 
atomic weight)  valence. The number of ions in an equivalent  
Avogardro’s number/valence. Avogadro’s number  6.02  1023. An 
equivalent contains 6.02  1023 electron charges or 96,500 coulombs, 
which is 1 faraday. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
56 3 SOIL MINERALOGY 
erals of pyrophyllite and talc may form. These 
minerals cleave parallel to the layers. 
2. In some minerals (e.g., kaolinite, brucite, gibb-site), 
there are opposing layers of oxygens and 
hydroxyls or hydroxyls and hydroxyls. Hydrogen 
bonding then develops between the layers as well 
as van der Waals bonding. Hydrogen bonds re-main 
stable in the presence of water. 
Copyrighted Material 
3. Neutral silicate layers that are separated by 
highly polar water molecules may be held to-gether 
by hydrogen bonds. 
4. Cations needed for electrical neutrality may be in 
positions that control interlayer bonding. In mi-cas, 
some of the silicon is replaced by aluminum 
in the silica sheets. The resulting charge defi-ciency 
is partly balanced by potassium ions be-tween 
the unit cell layers. The potassium ion just 
fits into the holes formed by the bases of the 
silica tetrahedra (Fig. 3.12). As a result, it gen-erates 
a strong bond between the layers. In the 
chlorites, the charge deficiencies from substitu-tions 
in the octahedral sheet of the 21 sandwich 
are balanced by excess charge on the single-sheet 
layer interleaved between the three-sheet layers. 
This provides a strongly bonded structure that 
while exhibiting cleavage will not separate in the 
presence of water or other polar liquids. 
5. When the surface charge density is moderate, as 
in smectite and vermiculite, the silicate layers 
readily adsorb polar molecules, and also the ad-sorbed 
cations may hydrate, resulting in layer 
separation and expansion. The strength of the in-terlayer 
bond is low and is a strong function of 
charge distribution, ion hydration energy, surface 
ion configuration, and structure of the polar mol-ecule. 
Smectite and vermiculite particles adsorb water be-tween 
the unit layers and swell, whereas particles of 
the nonclay minerals, pyrophyllite and talc, which have 
comparable structures, do not. There are two possible 
reasons (van Olphen, 1977): 
1. The interlayer cations in smectite hydrate, and 
the hydration energy overcomes the attractive 
forces between the unit layers. There are no in-terlayer 
cations in pyrophyllite; hence, no swell-ing. 
2. Water does not hydrate the cations but is ad-sorbed 
on oxygen surfaces by hydrogen bonds. 
There is no swelling in pyrophyllite and talc be-cause 
the surface hydration energy is too small 
to overcome the van der Waals forces between 
layers, which are greater in these minerals be-cause 
of a smaller interlayer distance. 
Whatever the reason, the smectite minerals are the 
dominant source of swelling in the expansive soils that 
are so prevalent throughout the world. 
3.16 THE 11 MINERALS 
The kaolinite–serpentine minerals are composed of al-ternating 
silica and octahedral sheets as shown sche-matically 
in Fig. 3.16. The tips of the silica tetrahedra 
and one of the planes of atoms in the octahedral sheet 
are common. The tips of the tetrahedra all point in the 
same direction, toward the center of the unit layer. In 
the plane of atoms common to both sheets, two-thirds 
of the atoms are oxygens and are shared by both sili-con 
and the octahedral cations. The remaining atoms 
in this plane are (OH) located so that each is directly 
below the hole in the hexagonal net formed by the 
bases of the silica tetrahedra. If the octahedral layer is 
brucite, then a mineral of the serpentine subgroup re-sults, 
whereas dioctahedral gibbsite layers give clay 
minerals in the kaolinite subgroup. Trioctahedral 11 
minerals are relatively rare, usually occur mixed with 
kaolinite or illite, and are hard to identify. A diagram-matic 
sketch of the kaolinite structure is shown in Fig. 
3.17. The structural formula is (OH)8Si4Al4O10, and the 
charge distribution is indicated in Fig. 3.18. 
Mineral particles of the kaolinite subgroup consist 
of the basic units stacked in the c direction. The bond-ing 
between successive layers is by both van der Waals 
forces and hydrogen bonds. The bonding is sufficiently 
strong that there is no interlayer swelling in the pres-ence 
of water. 
Because of slight differences in the oxygen-to-oxygen 
distances in the tetrahedral and octahedral lay-ers, 
there is some distortion of the ideal tetrahedral 
network. As a result, kaolinite, which is the most abun-dant 
member of the subgroup and a common soil min-eral, 
is triclinic instead of monoclinic. The unit cell 
dimensions are a  5.16 A˚ 
, b  8.94 A˚ 
, c  7.37 A˚ 
, 
  91.8,   104.5, and 	  90. 
Variations in stacking of layers above each other, 
and possibly in the position of aluminum ions within 
the available sites in the octahedral sheet, produce dif-ferent 
members of the kaolinite subgroup. The dickite 
unit cell is made up of two unit layers, and the nacrite 
unit cell contains six. Both appear to be formed by 
hydrothermal processes. Dickite is fairly common as 
secondary clay in the pores of sandstone and in coal 
beds. Neither dickite nor nacrite is common in soils. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
THE 11 MINERALS 57 
Figure 3.16 Schematic diagrams of the structures ˚ 
Aof kaolinite and serpentine: (a) kaolinite 
and (b) serpentine. 
Material 
Copyrighted is about 10.1 . The difference between these 
Figure 3.17 Diagrammatic sketch of the kaolinite structure. 
Halloysite 
Halloysite is a particularly interesting mineral of the 
kaolinite subgroup. Two distinct endpoint forms of this 
mineral exist, as shown in Fig. 3.19; one, a hydrated 
form consisting of unit kaolinite layers separated from 
each other by a single layer of water molecules and 
having the composition (OH)8Si4Al4O10  4H2O, and 
the other, a nonhydrated form having the same unit 
layer structure and chemical composition as kaolinite. 
The basal spacing in the c direction d(001) for the non-hydrated 
form is about 7.2 A˚ 
, as for kaolinite. Because 
of the interleaved water layer, d(001) for hydrated hal-loysite 
Figure 3.18 Charge distribution on kaolinite. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
58 3 SOIL MINERALOGY 
Figure 3.19 Schematic diagrams of the structure of halloysite: (a) halloysite (10 ) and (b) 
halloysite (7 ). 
Material 
Copyrighted varieties. 
˚ 
A˚ 
AFigure 3.20 Electron photomicrograph of well-crystallized 
kaolinite from St. Austell, Cornwall, England. Picture width 
is 17 m (Tovey, 1971). 
values, 2.9 A˚ 
, is the approximate thickness of a single 
layer of water molecules. 
The recommended terms for the two forms of hal-loysite 
are halloysite (7 A˚ 
) and halloysite (10 A˚ 
). 
Transformation from halloysite (10 A˚ 
) to halloysite (7 
A˚ 
) by dehydration can occur at relatively low temper-atures 
and is irreversible. Halloysite is often found in 
soils formed from volcanic parent materials in wet en-vironments. 
It can be responsible for special properties 
and problems in earthwork construction, as discussed 
later in this book. 
Isomorphous Substitution and Exchange Capacity 
Whether or not measurable isomorphous substitution 
exists within the structure of the kaolinite minerals is 
uncertain. Nevertheless, values of cation exchange ca-pacity 
in the range of 3 to 15 meq/100 g for kaolinite 
and from 5 to 40 meq/100 g for halloysite have been 
measured. Thus, kaolinite particles possess a net neg-ative 
charge. Possible sources are: 
1. Substitution of Al3 for Si4 in the silica sheet or 
a divalent ion for Al3 in the octahedral sheet. 
Replacement of only 1 Si in every 400 would be 
adequate to account for the exchange capacity. 
2. The hydrogen of exposed hydroxyls may be re-placed 
by exchangeable cations. According to 
Grim (1968), however, this mechanism is not 
likely because the hydrogen would probably not 
be replaceable under the conditions of most 
exchange reactions. 
3. Broken bonds around particle edges may give un-satisfied 
charges that are balanced by adsorbed 
cations. 
Kaolinite particles are charged positively on their 
edges when in a low pH (acid) environment, but neg-atively 
charged in a high pH (basic) environment. Low 
exchange capacities are measured under low pH con-ditions 
and high exchange capacities are obtained for 
determinations at high pH. This suggests that broken 
bonds are at least a partial source of exchange capacity. 
That a positive cation exchange capacity is measured 
under low pH conditions when edges are positively 
charged indicates that some isomorphous substitution 
must exist also. 
As interlayer separation does not occur in kaolinite, 
balancing cations must adsorb on the exterior surfaces 
and edges of the particles. 
Morphology and Surface Area 
Well-crystallized particles of kaolinite (Fig. 3.20), na-crite, 
and dickite occur as well-formed six-sided plates. 
The lateral dimensions of these plates range from 
about 0.1 to 4 m, and their thicknesses are from about 
0.05 to 2 m. Poorly crystallized kaolinite generally 
occurs as less distinct hexagonal plates, and the parti-cle 
size is usually smaller than for the well-crystallized 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
SMECTITE MINERALS 59 
Halloysite (10 A˚ 
) occurs as cylindrical tubes of 
overlapping sheets of the kaolinite type (Fig. 3.21). 
The c axis at any point nearly coincides with the tube 
radius. The formation of tubes has been attributed to a 
misfit in the b direction of the silica and gibbsite sheets 
(Bates et al., 1950). The b dimension in kaolinite is 
8.93 A˚ 
; in gibbsite it is only 8.62 A˚ 
. This means that 
the (OH) spacing in gibbsite sheets is stretched in order 
to obtain a fit with the silica sheet. Evidently, in hal-loysite 
Copyrighted Material 
(10 A˚ 
), the reduced interlayer bond, caused by 
the intervening layer of water molecules, enables the 
(OH) layer to revert to 8.62 A˚ 
, resulting in a curvature 
with the hydroxyls on the inside and the bases of the 
silica tetrahedra on the outside. The outside diameters 
of the tubular particles range from about 0.05 to 0.20 
m, with a median value of 0.07 m. The wall thick-ness 
is about 0.02 m. The tubes range in length from 
a fraction of a micrometer to several micrometers. Dry-ing 
of halloysite (10 A˚ 
) may result in splitting or un-rolling 
of the tubes. The specific surface area of 
kaolinite is about 10 to 20 m2 /g of dry clay; that of 
halloysite (10 A˚ 
) is 35 to 70 m2/g. 
3.17 SMECTITE MINERALS 
Structure 
The minerals of the smectite group have a prototype 
structure similar to that of pyrophyllite, consisting of 
an octahedral sheet sandwiched between two silica 
sheets, as shown schematically in Fig. 3.22 and dia-grammatically 
in three dimensions in Fig. 3.23. All the 
tips of the tetrahedra point toward the center of the 
unit cell. The oxygens forming the tips of the tetra-hedra 
are common to the octahedral sheet as well. The 
anions in the octahedral sheet that fall directly above 
Figure 3.21 Electron photomicrograph of halloysite from 
Bedford, Indiana. Picture width is 2 m (Tovey, 1971). 
and below the hexagonal holes formed by the bases of 
the silica tetrahedra are hydroxyls. 
The layers formed in this way are continuous in the 
a and b directions and stacked one above the other in 
the c direction. Bonding between successive layers is 
by van der Waals forces and by cations that balance 
charge deficiencies in the structure. These bonds are 
weak and easily separated by cleavage or adsorption 
of water or other polar liquids. The basal spacing in 
the c direction, d(001), is variable, ranging from about 
9.6 A˚ 
to complete separation. 
The theoretical composition in the absence of 
isomorphous substitutions is (OH)4Si8Al4O20  
n(interlayer)H2O. The structural configuration and cor-responding 
charge distribution are shown in Fig. 3.24. 
The structure shown is electrically neutral, and the 
atomic configuration is essentially the same as that in 
the nonclay mineral pyrophyllite. 
Isomorphous Substitution in the Smectite Minerals 
Smectite minerals differ from pyrophyllite in that there 
is extensive isomorphous substitution for silicon and 
aluminum by other cations. Aluminum in the octahe-dral 
sheet may be replaced by magnesium, iron, zinc, 
nickel, lithium, or other cations. Aluminum may re-place 
up to 15 percent of the silicon ions in the tetra-hedral 
sheet. Possibly some of the silicon positions can 
be occupied by phosphorous (Grim, 1968). 
Substitutions for aluminum in the octahedral sheet 
may be one-for-one or three-for-two (aluminum oc-cupies 
only two-thirds of the available octahedral sites) 
in any combination from a few to complete replace-ment. 
The resulting structure, however, is either almost 
exactly dioctahedral (montmorillonite subgroup) or 
trioctahedral (saponite subgroup). The charge defi-ciency 
resulting from these substitutions ranges from 
0.5 to 1.2 per unit cell. Usually, it is close to 0.66 per 
unit cell. A charge deficiency of this amount would 
result from replacement of every sixth aluminum by a 
magnesium ion. Montmorillonite, the most common 
mineral of the group, has this composition. Charge de-ficiencies 
that result from isomorphous substitution are 
balanced by exchangeable cations located between the 
unit cell layers and on the surfaces of particles. 
Some minerals of the smectite group and their com-positions 
are listed in Table 3.5. An arrow indicates the 
source of the charge deficiency, which has been as-sumed 
to be 0.66 per unit cell in each case. Sodium is 
indicated as the balancing cation. The formulas should 
be considered indicative of the general character of the 
mineral, but not as absolute, because a variety of com-positions 
can exist within the same basic crystal struc-ture. 
Because of the large amount of unbalanced 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
60 3 SOIL MINERALOGY 
Figure 3.22 Schematic diagrams of the structures of the smectite minerals: (a) montmoril-lonite 
and (b) saponite. 
Material 
Copyrighted Figure 3.23 Diagrammatic sketch of the montmorillonite 
structure. 
Figure 3.24 Charge distribution in pyrophyllite (type struc-ture 
for montmorillonite). 
substitution in the smectite minerals, they have high 
cation exchange capacities, generally in the range of 
80 to 150 meq/100 g. 
Morphology and Surface Area 
Montmorillonite may occur as equidimensional flakes 
that are so thin as to appear more like films, as shown 
in Fig. 3.25. Particles range in thickness from 1-nm 
unit layers upward to about 1/100 of the width. The 
long axis of the particle is usually less than 1 or 2 m. 
When there is a large amount of substitution of iron 
and/or magnesium for aluminum, the particles may be 
lath or needle shaped because the larger Mg2 and Fe3 
ions cause a directional strain in the structure. 
The specific surface area of smectite can be very 
large. The primary surface area, that is, the surface area 
exclusive of interlayer zones, ranges from 50 to 120 
m2 /g. The secondary specific surface that is exposed 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
SMECTITE MINERALS 61 
Table 3.5 Some Minerals of the Smectite Group 
Mineral 
Tetrahedral Sheet 
Substitutions 
Octahedral Sheet 
Substitutions Formula/Unit Cella 
Dioctahedral, Smectites or 
Montmorillonites 
Montmorillonite None 1Mg2 for every sixth Al3 (OH)4Si8(Al3.34Mg0.66) O20 
↓ 
Na0.66 
Copyrighted Material 
Beidellite Al for Si None (OH)4(Si6.34Al1.66) Al4.34O20 
↓ 
Na0.66 
Nontronite Al for Si Fe3 for Al (OH)4(Si7.34Al0.66) Fe4 
3O20 
↓ 
Na0.66 
Trioctahedral, Smectites, 
or Saponites 
Hectorite None Li for Mg (OH)4Si8(Mg5.34Li0.66) O20 
↓ 
Na0.66 
Saponite Al for Si Fe3 for Mg (OH)4(Si7.34Al0.66) Mg6O20 
↓ 
Na0.66 
Sauconite Al for Si Zn for Mg (OH)4(Si8yAly)(Zn6xMgx) O20 
↓ 
Na0.66 
aTwo formula units are needed to give one unit cell. 
After Ross and Hendricks (1945); Marshall (1964); and Warshaw and Roy (1961). 
Figure 3.25 Electron photomicrograph of montmorillonite 
(bentonite) from Clay Spur, Wyoming. Picture width is 7.5 
m (Tovey, 1971). 
by expanding the lattice so that polar molecules can 
penetrate between layers can be up to 840 m2/g. 
Bentonite 
A very highly plastic, swelling clay material known as 
bentonite is very widely used for a variety of purposes, 
ranging from drilling mud and slurry walls to clarifi-cation 
of beer and wine. The bentonite familiar to most 
geoengineers is a highly colloidal, expansive alteration 
product of volcanic ash. It has a liquid limit of 500 
percent or more. It is widely used as a backfill during 
the construction of slurry trench walls, as a soil ad-mixture 
for construction of seepage barriers, as a grout 
material, as a sealant for piezometer installations, and 
for other special applications. 
When present as a major constituent in soft shale or 
as a seam in rock formations, bentonite may be a cause 
of continuing slope stability problems. Slide problems 
at Portuguese Bend along the Pacific Ocean in southern 
California, in the Bearpaw shale in Saskatchewan, and 
in the Pierre shale in South Dakota are in large mea- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
62 3 SOIL MINERALOGY 
sure due to the high content of bentonite. Stability 
problems in underground construction may be caused 
by the presence of montmorillonite in joints and faults 
(Brekke and Selmer-Olsen, 1965). 
3.18 MICALIKE CLAY MINERALS 
Illite is the most commonly found clay mineral in soils 
encountered in engineering practice. Its structure is 
quite similar to that of muscovite mica, and it is some-times 
3.26b. The actual thickness of the water layer depends 
on the cations that balance the charge deficiencies in 
Copyrighted Material 
referred to as hydrous mica. Vermiculite is also 
often found as a clay phase constituent of soils. Its 
structure is related to that of biotite mica. 
Structure 
The basic structural unit for the muscovite (white mica) 
is shown schematically in Fig. 3.26a. It is the three-layer 
silica–gibbsite–silica sandwich that forms pyro-phyllite, 
with the tips of all the tetrahedra pointing 
toward the center and common with octahedral sheet 
ions. Muscovite differs from pyrophyllite, however, in 
that about one-fourth of the silicon positions are filled 
by aluminum, and the resulting charge deficiency is 
balanced by potassium between the layers. The layers 
are continuous in the a and b directions and stacked in 
the c direction. The radius of the potassium ion, 1.33 
, is such that it fits snugly in the 1.32 A˚ 
radius hole 
formed by the bases of the silica tetrahedra. It is in 12- 
fold coordination with the 6 oxygens in each layer. 
A diagrammatic three-dimensional sketch of the 
muscovite structure is shown in Fig. 3.27. The struc-tural 
Figure 3.26 Schematic diagram of the structures of muscovite, illite, and vermiculite: (a) 
muscovite and illite and (b) vermiculite. 
A˚ 
configuration and charge distribution are shown 
in Fig. 3.28. The unit cell is electrically neutral and 
has the formula (OH)4K2(Si6Al2)Al4O20. Muscovite is 
the dioctahedral end member of the micas and contains 
only Al3 in the octahedral layer. Phlogopite (brown 
mica) is the trioctahedral end member, with the octa-hedral 
positions filled entirely by magnesium. It has 
the formula (OH)4K2(Si6Al2)Mg6O20. Biotite (black 
mica) is trioctahedral, with the octahedral positions 
filled mostly by magnesium and iron. It has the general 
formula (OH)4K2(Si6Al2)(MgFe)6O20. The relative pro-portions 
of magnesium and iron may vary widely. 
Illite differs from mica in the following ways (Grim, 
1968): 
1. Fewer of the Si4 positions are filled by Al3 in 
illite. 
2. There is some randomness in the stacking of lay-ers 
in illite. 
3. There is less potassium in illite. Well-organized 
illite contains 9 to 10 percent K2O (Weaver and 
Pollard, 1973). 
4. Illite particles are much smaller than mica parti-cles. 
Some illite may contain magnesium and iron in the 
octahedral sheet as well as aluminum (Marshall, 1964). 
Iron-rich illite, usually occurring as earthy green pel-lets, 
is termed glauconite. 
The vermiculite structure consists of regular inter-stratification 
of biotite mica layers and double molec-ular 
layers of water, as shown schematically in Fig. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
MICALIKE CLAY MINERALS 63 
Material 
Copyrighted Figure 3.27 Diagrammatic sketch of the structure of muscovite. 
Figure 3.28 Charge distribution in muscovite. 
the biotitelike layers. With magnesium or calcium 
present, which is the usual case in nature, there are 
two water layers, giving a basal spacing of 14 A˚ 
. A 
general formula for vermiculite is 
(OH)4(MgCa)x(Si8
xAlx)(MgFe)6O20yH2O 
x  1 to 1.4 y  8 
Isomorphous Substitution and Exchange Capacity 
There is extensive isomorphous substitution in illite 
and vermiculite. The charge deficiency in illite is 1.3 
to 1.5 per unit cell. It is located primarily in the silica 
sheets and is balanced partly by the nonexchangeable 
potassium between layers. Thus, the cation exchange 
capacity of illite is less than that of smectite, amount-ing 
to 10 to 40 meq/100 g. Values greater than 10 to 
15 meq/100 g may be indicative of some expanding 
layers (Weaver and Pollard, 1973). In the absence of 
fixed potassium the exchange capacity would be about 
150 meq/100 g. Interlayer bonding by potassium is so 
strong that the basal spacing of illite remains fixed at 
10 A˚ 
in the presence of polar liquids. 
The charge deficiency in vermiculite is 1 to 1.4 per 
unit cell. Since the interlayer cations are exchangeable, 
the exchange capacity of vermiculite is high, amount- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
64 3 SOIL MINERALOGY 
Copyrighted Material 
Figure 3.30 Schematic diagram of the structure of chlorite. 
ing to 100 to 150 meq/100 g. The basal spacing, d(001), 
is influenced by both the type of cation and the hy-dration 
state. With potassium or ammonium in the 
exchange positions, the basal spacing is only 10.5 to 
11 A˚ 
. Lithium gives 12.2 A˚ 
. Interlayer water can be 
driven off by heating to temperatures above 100C. 
This dehydration is accompanied by a reduction in 
basal spacing to about 10 A˚ 
. The mineral quickly re-hydrates 
and expands again to 14 A˚ 
when exposed to 
moist air at room temperature. 
Morphology and Surface Area 
Illite usually occurs as very small, flaky particles 
mixed with other clay and nonclay materials. High-purity 
deposits of illite are uncommon. The flaky par-ticles 
may have a hexagonal outline if well crystallized. 
The long axis dimension ranges from 0.1 m or less 
to several micrometers, and the plate thickness may be 
as small as 3 nm. An electron photomicrograph of illite 
is shown in Fig. 3.29. Vermiculite may occur in nature 
as large crystalline masses having a sheet structure 
somewhat similar in appearance to mica. In soils, ver-miculite 
occurs as small particles mixed with other 
clay minerals. 
The specific surface area of illite is about 65 to 100 
m2 /g. The primary surface of vermiculites is 40 to 80 
m2 /g, and the secondary (interlayer) surface may be as 
high as 870 m2/g. 
3.19 OTHER CLAY MINERALS 
Chlorite Minerals 
Structure The chlorite structure consists of alter-nating 
micalike and brucitelike layers as shown sche-matically 
in Fig. 3.30. The structure is similar to that 
Figure 3.29 Electron photomicrograph of illite from Morris, 
Illinois. Picture width is 7.5 m (Tovey, 1971). 
of vermiculite, except that an organized octahedral 
sheet replaces the double water layer between mica 
layers. The layers are continuous in the a and b direc-tions 
and stacked in the c direction. The basal spacing 
is fixed at 14 A˚ 
. 
Isomorphous Substitution The central sheet of the 
mica layer is trioctahedral, with magnesium as the pre-dominant 
cation. There is often partial replacement of 
Mg2 by Al3, Fe2 and Fe3. There is substitution of 
Al3 for Mg2 in the brucitelike layer. The various 
members of the chlorite group differ in the kind and 
amounts of substitution and in the stacking of succes-sive 
layers. The cation exchange capacity of chlorites 
is in the range of 10 to 40 meq/100 g. 
Morphology Chlorite minerals occur as micro-scopic 
grains of platy morphology and poorly defined 
crystal edges in altered igneous and metamorphic rocks 
and their derived soils. In soils, chlorites always appear 
to occur in mixtures with other clay minerals. 
Chain Structure Clay Minerals 
A few clay minerals are formed from bands (double 
chains) of silica tetrahedra. These include attapulgite 
and imogolite. They have lathlike or fine threadlike 
morphologies, with particle diameters of 5 to 10 nm 
and lengths up to 4 to 5 m. An electron photomicro-graph 
of bundles of attapulgite particles is shown in 
Fig. 3.31. 
Although these minerals are not frequently encoun-tered, 
attapulgite is commercially mined and is used as 
a drilling mud in saline and other special environments 
because of its high stability in suspensions. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
DETERMINATION OF SOIL COMPOSITION 65 
Copyrighted Material 
Figure 3.31 Electron photomicrograph of attapulgite from 
Attapulgis, Georgia. Picture width is 4.7 m (Tovey, 1971). 
Mixed-Layer Clays 
More than one type of clay mineral is usually found 
in most soils. Because of the great similarity in crystal 
structure among the different minerals, interstratifica-tion 
of two or more layer types often occurs within a 
single particle. Interstratification may be regular, with 
a definite repetition of the different layers in sequence, 
or it may be random. According to Weaver and Pollard 
(1973), randomly interstratified clay minerals are sec-ond 
only to illite in abundance. 
The most abundant mixed-layer material is com-posed 
of expanded water-bearing layers and contracted 
non-water-bearing layers. Montmorillonite–illite is 
most common, and chlorite–vermiculite and chlorite– 
montmorillonite are often found. Rectorite is an inter-stratified 
clay with high charge, micalike layers with 
fixed interlayer cations alternating in a regular manner 
with low-charge montmorillonite-like layers containing 
exchangeable cations capable of hydration. 
Noncrystalline Clay Materials 
Allophane Clay materials that are so poorly crys-talline 
that a definite structure cannot be determined 
are termed allophane. Such material is amorphous to 
X-rays because there is insufficient long-range order of 
the octahedral and tetrahedral units to produce sharp 
diffraction effects, although in some cases there may 
be diffraction bands. Allophane has no definite com-position 
or shape and may exhibit a wide range of 
physical properties. Some noncrystalline clay material 
is probably contained in all fine-grained soils. It is 
common in volcanic soils because of the abundance of 
glass particles. 
Oxides All soils probably contain some amount of 
colloidal oxides and hydrous oxides (Marshall, 1964). 
The oxides and hydroxides of aluminum, silicon, and 
iron are most frequently found. These materials may 
occur as gels or precipitates and coat mineral particles, 
or they may cement particles together. They may also 
occur as distinct crystalline units; for example, gibb-site, 
boehmite, hematite, and magnetite. Limonite and 
bauxite, which are noncrystalline mixtures of iron and 
aluminum hydroxides, are also sometimes found. 
Oxides are particularly common in soils formed 
from volcanic ash and in tropical residual soils. Some 
soils rich in allophane and oxides may exhibit signif-icant 
irreversible decreases in plasticity and increases 
in strength when dried. Many are susceptible to break-down 
and strength loss when subjected to traffic or 
manipulation during earthwork construction (Mitchell 
and Sitar, 1982; Mitchell and Coutinho, 1991). 
3.20 SUMMARY OF CLAY MINERAL 
CHARACTERISTICS 
The important structural, compositional, and morpho-logical 
characteristics of the important clay minerals 
are summarized in Table 3.6. Data on the structural 
characteristics of the tetrahedral and octahedral sheet 
structures are included. 
3.21 DETERMINATION OF SOIL 
COMPOSITION 
Introduction 
Identification of the fine-grained minerals in a soil is 
usually done by X-ray diffraction. Simple chemical 
tests can be used to indicate the presence of organic 
matter and other constituents. The microscope may be 
used to identify the constituents of the nonclay frac-tion. 
Accurate determination of the proportions of dif-ferent 
mineral, organic, and amorphous solid material 
in a soil, while probably possible with the expenditure 
of great time and at great cost, is unlikely to be worth-while 
owing to our inability to make exact quantitative 
links from composition to properties. Accordingly, 
from knowledge of grain size distribution, the relative 
intensities of different X-ray diffraction peaks, and a 
few other simple tests a semiquantitative analysis may 
be made that is usually adequate for most purposes. 
A general approach is given in this section for the 
determination of soil composition, some of the tech-niques 
are described briefly, and criteria for identifi-cation 
of important soil constituents are stated. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
66 3 SOIL MINERALOGY 
Table 3.6 Summary of Clay Mineral Characteristics 
Structural 
1. Silica Tetrahedron: Si atom at center. Tetrahedron units form hexagonal network  Si4O8(OH)4 
2. Gibbsite Sheet: Aluminum in octahedral coordination. Two-thirds of possible positions filled. Al2(OH)—O—O  2.60 A˚ 
. 
3. Brucite Sheet: Magnesium in octahedral coordination. All possible positions filled. Mg2(OH)—O—O  2.60 A˚ 
. 
Type 
Subgroup and 
Schematic Structure Mineral Complete Formula / Unit Cella 
Octahedral Layer 
Cations 
Tetrahedral Layer 
Cations 
Structure 
Isomorphous Substitution Interlayer Bond 
Allophane Allophanes Amorphous — — 
Copyrighted Material 
Kaolinite Kaolinite (OH)8Si4Al4O11 Al4 Si4 Little O—OH 
Hydrogen Strong 
11 
Dickite 
Nacrite 
Halloysite 
(dehydrated) 
Halloysite 
(hydrated) 
(OH)8Si4Al4O10 
(OH)8Si4Al4O10 
(OH)8Si4Al4O10 
(OH)8Si4Al4O10  4H2O 
Al4 
Al4 
Al4 
Al4 
Si4 
Si4 
Si4 
Si4 
Little 
Little 
Little 
Little 
O—OH 
Hydrogen Strong 
O—OH 
Hydrogen Strong 
O—OH 
Hydrogen Strong 
O—OH 
Hydrogen Strong 
Montmorillonite 
(OH)4Si8Al4O20  NH2O 
(Theoretical 
Unsubsitituted) 
Montmorillonite (OH)4Si8(Al3.34Mg.66O20nH2O 
↓ * 
Na.66 
Al3.34Mg.66 Si8 Mg for Al, Net charge 
always  0.66- / unit 
cell 
O—O 
Very weak 
expanding lattice 
Beidellite 
Nontronite 
(OH)4(Si7.34Al66)(Al4)O20nH2O 
↓ 
Na.66 
(OH)4(Si7.34Al.66)Fe4 
3O20nH2O 
↓ 
Na.66 
Al4 
Fe4 
Si7.34Al.66 
Si7.34Al.66 
Al for Si, Net charge 
always  0.66- / for 
unit cell 
Fe for Al, Al for Si, Net 
charge always  0.66- 
/ for unit cell 
O—O 
Very weak 
expanding lattice 
O—O 
Very weak 
expanding lattice 
21 Saponite Hectorite 
Saponite 
Sauconite 
(OH)4Si8(Mg5.34Li.66)P20nH2O 
↓ 
Na.66 
(OH)4(Si7.34Al.66)Mg6O20nH2O 
↓ 
Na.66 
(Si6.94Al1.06)Al.66Fe.34Mg.36Zn4.80O20(OH)4 
↓  nH2O 
Na.66 
Mg5.34Li.66 
Mg, Fe3 
Al.44Fe.34Mg.36Zn4.80 
Si8 
Si7.34Al.66 
Si6.94Al1.06 
Mg, Li for Al, Net 
charge always  0.66- 
/ unit cell 
Mg for Al, Al for Si, 
Net charge always  
0.66- / for unit cell 
Zn for Al 
O—O 
Very weak 
expanding lattice 
O—O 
Very weak 
expanding lattice 
O—O 
Very weak 
expanding lattice 
Hydrous Mica (Illite) Illites (K, H2O)2(Si)8(Al,Mg,Fe)4,6O20(OH)4 (Al,Mg,Fe)4-6 (Al,Si)8 Some Si always replaced 
by Al, Balanced by K 
between layers. 
K ions; strong 
Vermiculite Vermiculite (OH)4(Mg,Ca)x(Si8xAlx)(Mg.Fe)6O20.yH2O 
x  1 to 1.4, y  8 
(Mg,Fe)6 (Si,Al)8 Al for Si not charge of 1 
to 1.4 / unit cell 
Weak 
211 Chlorite Chlorite 
(Several varieties 
known) 
(OH)4(SiAl)8(Mg.Fe)6O20 (21 layer) 
(MgAl)6(OH)12 interlayer 
(Mg,Fe)6(21 layer) 
(Mg,Al)6 interlayer 
(Si,Al)8 Al for Si in 21 layer 
Al for Mg in interlayer 
Chain 
Structure 
Sepiolite 
Attapulgite 
Si4O11(Mg.H2)3H2O2(H2O) 
(OH2)4
(OH)2Mg5Si8O20.4H2O 
Fe or Al for Mg 
Some for Al for Si Weak  chains 
linked by 0 
a Arrows indicate source of charge deficiency. Equivalent Na listed as balancing cation. Two formula units (Table 3.4) are required per unit cell. 
b Electron microscope data. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
DETERMINATION OF SOIL COMPOSITION 67 
Table 3.6 (Continued) 
Units 
All bases in same plane. O—O  2.55 A˚ 
—Space for Si  0.55 A˚ 
—Thickness8 4.93 A˚ 
. C—C height  2.1 A˚ 
. 
OH—OH  2.94 A˚ 
. Space for ion  0.61 A˚ 
. Thickness of unit  5.05 A˚ 
. Dioctahedral. 
OH—OH  2.94 A˚ 
. Space for ion  0.61 A˚ 
. Thickness of unit  5.05 A˚ 
. Trioctahedral. 
Structure 
Crystal Structure Basal Spacing Shape Sizeb 
Cation Exchange 
Cap.(meq / 100 g) 
Specific 
Gravity 
Specific Surface 
m2 / g 
Occurrence in Soils 
of Engineering 
Interest 
Irregular, some-what 
rounded 
Material 
Copyrighted 0.05–1  
Common 
Triclinic 
a  5.14, b  8.93, c  7.37 
  91.6,   104.8, 	  89.9 
7.2 A˚ 
6-sided flakes 0.1–4  single 0.05–2  
to 3000  4000 
(stacks) 
3–15 2.60–2.68 10–20 Very common 
Monoclinic 
a  5.15, b  8.95, c  14.42 
  9648 
14.4 A˚ 
Unit cell 
contains 2 
unit layers 
6-sided flakes 0.07–300  2.5– 
1000  
1–30 Rare 
Almost Orthorhombic 
a  5.15, b  8.96, c  43 
  9020 
a  5.14 in O Plane 
a  5.06 in OH Plane 
b  8.93 in O Plane 
b  8.62 in OH Plane 
 layers curve 
43 A˚ 
7.2 A˚ 
10.1 A˚ 
Unit cell 
contains 6 
unit layers 
Random 
stacking of 
unit cells 
Water layer 
between unit 
cells 
Rounded flakes 
Tubes 
Tubes 
1   0.025– 
0.15  
0.07  O.D. 
0.04  I.D. 
1  long. 
5–10 
5–40 
2.55–2.56 
2.0–2.2 35–70 
Rare 
Occasional 
Occasional 
9.6A˚—Complete 
separation 
Dioctahedral Flakes (equi-dimensional) 
10 A˚ 
 up to 
10  
80–150 2.35–2.7 50–120 Primary 
700–840 Secondary 
Very common 
9.6A˚—Complete 
separation 
Dioctahedral Rare 
9.6A˚—Complete 
separation 
Dioctahedral Laths Breadth  1 / 5 
length to 
several   
unit cell 
110–150 2.2–2.7 Rare 
9.6A˚—Complete 
separation 
Trioctahedral To 1   unit 
cell breadth  
0.02  0.1 
17.5 Rare 
Trioctahedral Similar to 
mont. 
Similar to mont. 70–90 2.24–2.30 Rare 
Trioctahedral Brand laths 50 A˚ 
Thick Rare 
10 A˚ 
Both 
dioctrahedral 
and 
trioctahedral 
Flakes 0.003–0.1   
up to 10  
10–40 2.6–3.0 65–100 Very common 
a  5.34, b  9.20 
c  28.91,   9315 
10.5–14 A˚ 
Alternating 
Mica and 
double H2O 
layers 
Similar to illite 100–150 40–80 Primary 
870 Secondary 
Fairly common 
Monoclinic (Mainly) 
a  5.3, b  9.3 
c  28.52,   978 
14 A˚ 
Similar to illite 1  10–40 2.6–2.96 Common 
Monoclinic 
a  2  11.6, b  2  7.86 
c  5.33 
a0 Sin   12.9 b0  18 
c0  5.2 
Chain 
Double silica 
chains 
Flakes or fibers 
Laths Max, 4–5   
50–100 A˚ 
Width  2t 
20–30 
20–30 
2.08 Rare 
Occasional 
From Grim, R. E. (1968) Clay Mineralogy, 2d edition, McGraw-Hill, New York. Brown, G. (editor) (1961) The X-ray Identification and Crystal Structure of Clay Materials, Mineralogical 
Society (Clay Minerals Group), London. 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
68 3 SOIL MINERALOGY 
Methods for Compositional Analysis 
Methods and techniques that may be employed for de-termination 
of soil composition and study of soil grains 
include: 
1. Particle size analysis and separation 
2. Various pretreatments prior to mineralogical 
analysis 
Copyrighted Material 
3. Chemical analyses for free oxides, hydroxides, 
amorphous constituents, and organic matter 
4. Petrographic microscope study of silt and sand 
grains 
5. Electron microscope study 
6. X-ray diffraction for identification of crystalline 
minerals 
7. Thermal analysis 
8. Determination of specific surface area 
9. Chemical analysis for layer charge, cation 
exchange capacity, exchangeable cations, pH, 
and soluble salts 
10. Staining tests for identification of clays 
Procedures for determination of soil composition are 
described in detail in publications of the American So-ciety 
of Agronomy. Part 1—Physical and Mineralog-ical 
Methods provides a set of procedures for 
mineralogical analyses for use by soil scientists and 
engineers. Part 2—Microbiological and Biochemical 
Properties, published in 1994, is useful for determi-nations 
needed for bioremediation and other geoen-vironmental 
purposes. Part 3—Chemical Methods, 
published in 1996 contains methods for characterizing 
soil chemical properties as well as several methods for 
characterizing soil chemical processes. Part 4— 
Physical Methods, published in 2002, is an updated 
version of the physical methods covered in Part 1. For 
each method, principles are presented as well as the 
details of the method. In addition, the interpretation of 
results is discussed, and extensive bibliographies are 
given. 
Accuracy of Compositional Analysis 
Techniques for chemical analysis are generally of a 
high order of accuracy. However, this accuracy does 
not extend to the overall compositional analysis of a 
soil in terms of components of interest in understand-ing 
and quantifying behavior. This is because knowl-edge 
of the chemical composition of a soil is of limited 
value by itself. Chemical analysis of the solid phase of 
a soil does not indicate the organization of the ele-ments 
into crystalline and noncrystalline components. 
For quantitative mineralogical analysis of the clay 
fraction, it is usually necessary to assume that the 
properties of the mineral in the soil are the same as 
those of a reference mineral. However, different sam-ples 
of any given clay mineral may exhibit significant 
differences in composition, surface area, particle size 
and shape, and cation exchange capacity. Thus, selec-tion 
of ‘‘standard’’ minerals for reference is arbitrary. 
Quantitative clay mineral determinations cannot be 
made to an accuracy of more than about plus or minus 
a few percent without exhaustive chemical and min-eralogical 
tests. 
General Scheme for Compositional Analysis 
A general scheme for determination of the components 
of a soil is given in Fig. 3.32. Techniques of the most 
value for qualitative and semiquantitative analysis are 
indicated by a double asterisk, and those of particular 
use for explaining unusual properties are indicated by 
a single asterisk. The scheme shown is by no means 
the only one that could be used; a feedback approach 
is desirable wherein the results of each test are used to 
plan subsequent tests. Brief discussions of the various 
techniques listed in Fig. 3.32 are given below. X-ray 
diffraction analysis is treated in more detail in the next 
section because of its particular usefulness for the 
identification of fine-grained soil minerals. 
Grain Size Analysis Determination of particle size 
and size distribution is usually done using sieve anal-ysis 
for the coarse fraction [sizes greater than 74 m 
(i.e., 200 mesh sieve)] and by sedimentation methods 
for the fine fraction. Details of these methods are pre-sented 
in standard soil mechanics texts and in the stan-dards 
of the American Society for Testing and 
Materials (ASTM). Determination of sizes by sedi-mentation 
is based on the application of Stokes’s law 
for the settling velocity of spherical particles: 
	s  	w 2 v  D (3.2) 
18 
where 	s  unit weight of particle, 	w  unit weight 
of liquid,   viscosity of liquid, and D  diameter 
of sphere. Sizes determined by Stoke’s law are not ac-tual 
particle diameters but, rather, equivalent spherical 
diameters. Gravity sedimentation is limited to particle 
sizes in the range of about 0.2 mm to 0.2 m, the 
upper bound reflecting the size limit where flow around 
the particles is no longer laminar, and the lower bound 
representing a size where Brownian motion keeps par-ticles 
in suspension indefinitely. 
The times for particles of 2, 5, and 20 m equivalent 
spherical diameter to fall through water a distance of 
10 cm are about 8 h, 1.25 h, and 5 min, respectively, 
at 20C. At 30C the required times are about 6.5 h, 1 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
DETERMINATION OF SOIL COMPOSITION 69 
Material 
Copyrighted Figure 3.32 Flow sheet for compositional analysis of soils (adapted from Lambe and Martin, 
1954). 
h, and 4 min. A centrifuge can be used for accelerating 
the settlement of small particles and is the most prac-tical 
means for extracting particles smaller than about 
a micrometer in size. 
Sedimentation methods call for treatment of a soil– 
water suspension with a dispersing agent and thorough 
mixing prior to the start of the test. This causes break-down 
of aggregates of soil particles, and the degree of 
breakdown may vary greatly with the method of prep-aration. 
For example, the ASTM standard method of 
test permits the use of either an air dispersion cup or 
a blender-type mixer. The amount of material less than 
2 m equivalent spherical diameter may vary by as 
much as a factor of 2 by the two techniques. The re-lationship 
between the size distribution that results 
from laboratory preparation of the sample to that of 
the particles and aggregates in the natural soil is un-known. 
Optical and electron microscopes are sometimes 
used to study particle sizes and size distributions and 
to provide information on particle shape, aggregation, 
angularity, weathering, and surface texture. 
Pore Fluid Electrolyte The total concentration of 
soluble salts may be determined from the electrical 
conductivity of extracted pore fluid. Chemical or pho-tometric 
techniques may be used to determine the el-emental 
constituents of the extract (Rhoades, 1982). 
Removal of excess soluble salts by washing the sample 
with water or alcohol may be necessary before pro-ceeding 
with subsequent analysis. If they are not re- 
Copyright © 2005 John Wiley  Sons Retrieved from: www.knovel.com
63027 03
63027 03
63027 03
63027 03
63027 03
63027 03
63027 03
63027 03
63027 03
63027 03
63027 03
63027 03
63027 03

More Related Content

What's hot (19)

Soil colloidal chemistry
Soil colloidal chemistrySoil colloidal chemistry
Soil colloidal chemistry
 
The+clay+minerals+2007
The+clay+minerals+2007The+clay+minerals+2007
The+clay+minerals+2007
 
Clay mineral
Clay mineralClay mineral
Clay mineral
 
Clay mineralogy sivakugan
Clay mineralogy sivakuganClay mineralogy sivakugan
Clay mineralogy sivakugan
 
Cation exchange and it’s role on soil behaviour
Cation exchange and it’s role on soil behaviourCation exchange and it’s role on soil behaviour
Cation exchange and it’s role on soil behaviour
 
Soil Colloids
Soil ColloidsSoil Colloids
Soil Colloids
 
4 clay mineralogy
4 clay mineralogy4 clay mineralogy
4 clay mineralogy
 
Week 7 Clay And Ion Exchange (1)
Week 7 Clay And Ion Exchange (1)Week 7 Clay And Ion Exchange (1)
Week 7 Clay And Ion Exchange (1)
 
Preventing Clay Swelling
Preventing Clay Swelling Preventing Clay Swelling
Preventing Clay Swelling
 
Clay mineralogy sivakugan (Complete Soil Mech. Undestanding Pakage: ABHAY)
Clay mineralogy sivakugan (Complete Soil Mech. Undestanding Pakage: ABHAY)Clay mineralogy sivakugan (Complete Soil Mech. Undestanding Pakage: ABHAY)
Clay mineralogy sivakugan (Complete Soil Mech. Undestanding Pakage: ABHAY)
 
Silicate clay.(Morshed)
Silicate clay.(Morshed)Silicate clay.(Morshed)
Silicate clay.(Morshed)
 
Phyllosilicate clays
Phyllosilicate claysPhyllosilicate clays
Phyllosilicate clays
 
Clay mineral(mineralogy)
Clay mineral(mineralogy)Clay mineral(mineralogy)
Clay mineral(mineralogy)
 
Acquisition of K in Plants
Acquisition of K in PlantsAcquisition of K in Plants
Acquisition of K in Plants
 
1 introduction
1 introduction1 introduction
1 introduction
 
Phyllosilicate clays
Phyllosilicate claysPhyllosilicate clays
Phyllosilicate clays
 
Cec
CecCec
Cec
 
Soil Colloids: Properties, Nature, Types and Significance. sources of charges
Soil Colloids: Properties, Nature, Types and Significance. sources of chargesSoil Colloids: Properties, Nature, Types and Significance. sources of charges
Soil Colloids: Properties, Nature, Types and Significance. sources of charges
 
Clay
Clay  Clay
Clay
 

Similar to 63027 03

Advchemchapt10 101015122937-phpapp01
Advchemchapt10 101015122937-phpapp01Advchemchapt10 101015122937-phpapp01
Advchemchapt10 101015122937-phpapp01Cleophas Rwemera
 
Apchemunit10 111116094703-phpapp01
Apchemunit10 111116094703-phpapp01Apchemunit10 111116094703-phpapp01
Apchemunit10 111116094703-phpapp01Cleophas Rwemera
 
Structure of matter /certified fixed orthodontic courses by Indian dental aca...
Structure of matter /certified fixed orthodontic courses by Indian dental aca...Structure of matter /certified fixed orthodontic courses by Indian dental aca...
Structure of matter /certified fixed orthodontic courses by Indian dental aca...Indian dental academy
 
Structures of solid
Structures of solidStructures of solid
Structures of solidArinah Alias
 
Introduction to Solid State Physics.ppt
Introduction to Solid State Physics.pptIntroduction to Solid State Physics.ppt
Introduction to Solid State Physics.pptHimanshuKumar899318
 
Metals, semiconductors and semiconductors
Metals, semiconductors and semiconductorsMetals, semiconductors and semiconductors
Metals, semiconductors and semiconductorsShivaji Burungale
 
Physical properties of dental materials
Physical properties of dental materialsPhysical properties of dental materials
Physical properties of dental materialsDr. KRITI TREHAN
 
solid state ncert class 11 text book pdf
solid state ncert class 11 text book pdfsolid state ncert class 11 text book pdf
solid state ncert class 11 text book pdfHrithikaV
 
Chapter1 150109004625-conversion-gate01
Chapter1 150109004625-conversion-gate01Chapter1 150109004625-conversion-gate01
Chapter1 150109004625-conversion-gate01Cleophas Rwemera
 
Chapter 1: Material Structure and Binary Alloy System
Chapter 1: Material Structure and Binary Alloy SystemChapter 1: Material Structure and Binary Alloy System
Chapter 1: Material Structure and Binary Alloy Systemsyar 2604
 
PROPERTIES OF MATERIALS FOR ENGINEERING APPLICATION 22.pdf
PROPERTIES OF MATERIALS FOR ENGINEERING APPLICATION 22.pdfPROPERTIES OF MATERIALS FOR ENGINEERING APPLICATION 22.pdf
PROPERTIES OF MATERIALS FOR ENGINEERING APPLICATION 22.pdfEmmanuelIsiayinekife
 
How do the atoms of metal and ceramic form solid structures Use che.pdf
How do the atoms of metal and ceramic form solid structures Use che.pdfHow do the atoms of metal and ceramic form solid structures Use che.pdf
How do the atoms of metal and ceramic form solid structures Use che.pdfarishmarketing21
 
engineering material atomic structure of atom Lecture2.
engineering material atomic structure of atom  Lecture2.engineering material atomic structure of atom  Lecture2.
engineering material atomic structure of atom Lecture2.muhammadawaisqaiser
 

Similar to 63027 03 (20)

Advchemchapt10 101015122937-phpapp01
Advchemchapt10 101015122937-phpapp01Advchemchapt10 101015122937-phpapp01
Advchemchapt10 101015122937-phpapp01
 
Apchemunit10 111116094703-phpapp01
Apchemunit10 111116094703-phpapp01Apchemunit10 111116094703-phpapp01
Apchemunit10 111116094703-phpapp01
 
Structure of matter /certified fixed orthodontic courses by Indian dental aca...
Structure of matter /certified fixed orthodontic courses by Indian dental aca...Structure of matter /certified fixed orthodontic courses by Indian dental aca...
Structure of matter /certified fixed orthodontic courses by Indian dental aca...
 
Structures of solid
Structures of solidStructures of solid
Structures of solid
 
Nanotechnology-5.docx
Nanotechnology-5.docxNanotechnology-5.docx
Nanotechnology-5.docx
 
Structure of Matter
Structure of MatterStructure of Matter
Structure of Matter
 
Introduction to Solid State Physics.ppt
Introduction to Solid State Physics.pptIntroduction to Solid State Physics.ppt
Introduction to Solid State Physics.ppt
 
Metals, semiconductors and semiconductors
Metals, semiconductors and semiconductorsMetals, semiconductors and semiconductors
Metals, semiconductors and semiconductors
 
Physical properties of dental materials
Physical properties of dental materialsPhysical properties of dental materials
Physical properties of dental materials
 
Ap chem unit 10
Ap chem unit 10Ap chem unit 10
Ap chem unit 10
 
Adv chem chapt 10
Adv chem chapt 10Adv chem chapt 10
Adv chem chapt 10
 
Adv chem chapt 10
Adv chem chapt 10Adv chem chapt 10
Adv chem chapt 10
 
Atomic bonding
Atomic bondingAtomic bonding
Atomic bonding
 
solid state ncert class 11 text book pdf
solid state ncert class 11 text book pdfsolid state ncert class 11 text book pdf
solid state ncert class 11 text book pdf
 
Chapter1 150109004625-conversion-gate01
Chapter1 150109004625-conversion-gate01Chapter1 150109004625-conversion-gate01
Chapter1 150109004625-conversion-gate01
 
Chapter 1: Material Structure and Binary Alloy System
Chapter 1: Material Structure and Binary Alloy SystemChapter 1: Material Structure and Binary Alloy System
Chapter 1: Material Structure and Binary Alloy System
 
BONDING.pptx
BONDING.pptxBONDING.pptx
BONDING.pptx
 
PROPERTIES OF MATERIALS FOR ENGINEERING APPLICATION 22.pdf
PROPERTIES OF MATERIALS FOR ENGINEERING APPLICATION 22.pdfPROPERTIES OF MATERIALS FOR ENGINEERING APPLICATION 22.pdf
PROPERTIES OF MATERIALS FOR ENGINEERING APPLICATION 22.pdf
 
How do the atoms of metal and ceramic form solid structures Use che.pdf
How do the atoms of metal and ceramic form solid structures Use che.pdfHow do the atoms of metal and ceramic form solid structures Use che.pdf
How do the atoms of metal and ceramic form solid structures Use che.pdf
 
engineering material atomic structure of atom Lecture2.
engineering material atomic structure of atom  Lecture2.engineering material atomic structure of atom  Lecture2.
engineering material atomic structure of atom Lecture2.
 

More from Vcoi Vit

2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri312eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31Vcoi Vit
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri142eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14Vcoi Vit
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri132eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13Vcoi Vit
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri052eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05Vcoi Vit
 
Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02Vcoi Vit
 
Giao trinh tong hop sv50
Giao trinh tong hop sv50Giao trinh tong hop sv50
Giao trinh tong hop sv50Vcoi Vit
 
Giao trinh tong hop sv49
Giao trinh tong hop sv49Giao trinh tong hop sv49
Giao trinh tong hop sv49Vcoi Vit
 
Giao trinh tong hop sv48
Giao trinh tong hop sv48Giao trinh tong hop sv48
Giao trinh tong hop sv48Vcoi Vit
 
Giao trinh tong hop sv47
Giao trinh tong hop sv47Giao trinh tong hop sv47
Giao trinh tong hop sv47Vcoi Vit
 
Giao trinh tong hop sv46
Giao trinh tong hop sv46Giao trinh tong hop sv46
Giao trinh tong hop sv46Vcoi Vit
 
Giao trinh tong hop sv45
Giao trinh tong hop sv45Giao trinh tong hop sv45
Giao trinh tong hop sv45Vcoi Vit
 
Giao trinh tong hop sv43
Giao trinh tong hop sv43Giao trinh tong hop sv43
Giao trinh tong hop sv43Vcoi Vit
 
Giao trinh tong hop sv41
Giao trinh tong hop sv41Giao trinh tong hop sv41
Giao trinh tong hop sv41Vcoi Vit
 

More from Vcoi Vit (20)

2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri312eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri31
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri142eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri14
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri132eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri13
 
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri052eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
2eae0e86 e7c6-431d-aa08-b8f89ba71921 giaotrinhkinhtechinhtri05
 
Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46Giao trinh nuoi_trong_thuy_san46
Giao trinh nuoi_trong_thuy_san46
 
Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28Giao trinh nuoi_trong_thuy_san28
Giao trinh nuoi_trong_thuy_san28
 
Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23Giao trinh nuoi_trong_thuy_san23
Giao trinh nuoi_trong_thuy_san23
 
Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22Giao trinh nuoi_trong_thuy_san22
Giao trinh nuoi_trong_thuy_san22
 
Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21Giao trinh nuoi_trong_thuy_san21
Giao trinh nuoi_trong_thuy_san21
 
Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16Giao trinh nuoi_trong_thuy_san16
Giao trinh nuoi_trong_thuy_san16
 
Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05Giao trinh nuoi_trong_thuy_san05
Giao trinh nuoi_trong_thuy_san05
 
Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02Giao trinh nuoi_trong_thuy_san02
Giao trinh nuoi_trong_thuy_san02
 
Giao trinh tong hop sv50
Giao trinh tong hop sv50Giao trinh tong hop sv50
Giao trinh tong hop sv50
 
Giao trinh tong hop sv49
Giao trinh tong hop sv49Giao trinh tong hop sv49
Giao trinh tong hop sv49
 
Giao trinh tong hop sv48
Giao trinh tong hop sv48Giao trinh tong hop sv48
Giao trinh tong hop sv48
 
Giao trinh tong hop sv47
Giao trinh tong hop sv47Giao trinh tong hop sv47
Giao trinh tong hop sv47
 
Giao trinh tong hop sv46
Giao trinh tong hop sv46Giao trinh tong hop sv46
Giao trinh tong hop sv46
 
Giao trinh tong hop sv45
Giao trinh tong hop sv45Giao trinh tong hop sv45
Giao trinh tong hop sv45
 
Giao trinh tong hop sv43
Giao trinh tong hop sv43Giao trinh tong hop sv43
Giao trinh tong hop sv43
 
Giao trinh tong hop sv41
Giao trinh tong hop sv41Giao trinh tong hop sv41
Giao trinh tong hop sv41
 

63027 03

  • 1. between clay and nonclay minerals is that the nonclays are composed primarily of bulky particles; whereas, the particles of most of the clay minerals are platy, and in a few cases they are needle shaped or tubular. The great range in soil particle sizes in relation to other particulate materials, electromagnetic wave lengths, and other size-dependent factors can be seen in Fig. 3.2. The liquid phase of most soil systems is composed of water containing various types and amounts of dissolved electrolytes. Organic compounds, both soluble and immiscible, are found in soils at sites Copyrighted Material 35 CHAPTER 3 Soil Mineralogy Figure 3.1 Particle size ranges in soils. 3.1 IMPORTANCE OF SOIL MINERALOGY IN GEOTECHNICAL ENGINEERING Soil is composed of solid particles, liquid, and gas and ranges from very soft, organic deposits through less compressible clays and sands to soft rock. The solid particles vary in size from large boulders to minute particles that are visible only with the aid of the elec-tron microscope. Particle shapes range from nearly spherical, bulky grains to thin, flat plates and long, slender needles. Some organic material and noncrys-talline inorganic components are found in most natural fine-grained soils. A soil may contain virtually any el-ement contained in Earth’s crust; however, by far the most abundant are oxygen, silicon, hydrogen, and alu-minum. These elements, along with calcium, sodium, potassium, magnesium, and carbon, comprise over 99 percent of the solid mass of soils worldwide. Atoms of these elements are organized into various crystalline forms to yield the common minerals found in soil. Crystalline minerals comprise the greatest proportion of most soils encountered in engineering practice, and the amount of nonclay material usually exceeds the amount of clay. Nonetheless, clay and organic matter in a soil usually influence properties in a manner far greater than their abundance. Mineralogy is the primary factor controlling the size, shape, and properties of soil particles. These same fac-tors determine the possible ranges of physical and chemical properties of any given soil; therefore, a priori knowledge of what minerals are in a soil pro-vides intuitive insight as to its behavior. Commonly defined particle size ranges are shown in Fig. 3.1. The divisions between gravel, sand, silt, and clay sizes are arbitrary but convenient. Particles smaller than about 200 mesh sieve size (0.074 mm), which is the bound-ary between sand and silt sizes, cannot be seen by the naked eye. Clay can refer both to a size and to a class of minerals. As a size term, it refers to all constituents of a soil smaller than a particular size, usually 0.002 mm (2 m) in engineering classifications. As a mineral term, it refers to specific clay minerals that are distin-guished by (1) small particle size, (2) a net negative electrical charge, (3) plasticity when mixed with water, and (4) high weathering resistance. Clay minerals are primarily hydrous aluminum silicates. Not all clay par-ticles are smaller than 2 m, and not all nonclay par-ticles are coarser than 2 m; however, the amount of clay mineral in a soil is often closely approximated by the amount of material finer than 2 m. Thus, it is useful to use the terms clay size and clay mineral con-tent to avoid confusion. A further important difference Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 2. 36 Copyrighted Material Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 3. 37 Copyrighted Material Figure 3.2 Characteristics of particles and particle dispersoids (adapted from Stanford Re-search Institute Journal, Third Quarter, 1961). Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 4. 38 3 SOIL MINERALOGY that have been affected by chemical spills, leaking wastes, and contaminated groundwater. The gas phase, in partially saturated soils, is usually air, although or-ganic gases may be present in zones of high biological activity or in chemically contaminated soils. The mechanical properties of soils depend directly on interactions of these phases with each other and with applied potentials (e.g., stress, hydraulic head, electrical potential, and temperature). Because of these interactions, we cannot understand soil behavior in terms of the solid particles alone. Nonetheless, the structure of these particles tells us a great deal about their surface characteristics and their potential inter-actions 3.3 INTERATOMIC BONDING Primary Bonds Only the outer shell or valence electrons participate in the formation of primary interatomic bonds. There are Copyrighted Material Figure 3.3 Simplified representation of an atom. with adjacent phases. Interatomic and intermolecular bonding forces hold matter together. Unbalanced forces exist at phase boundaries. The nature and magnitude of these forces influence the formation of soil minerals, the structure, size, and shape of soil particles, and the physicochem-ical phenomena that determine engineering properties and behavior. In this chapter some aspects of atomic and intermolecular forces, crystal structure, structure stability, and characteristics of surfaces that are perti-nent to the understanding of soil behavior are sum-marized simply and briefly. This is followed by a somewhat more detailed treatment of soil minerals and their characteristics. 3.2 ATOMIC STRUCTURE Current concepts of atomic structure and interparticle bonding forces are based on quantum mechanics. An electron can have only certain values of energy. Elec-tronic energy can jump to a higher level by the ab-sorption of radiant energy or drop to a lower level by the emission of radiant energy. No more than two elec-trons in an atom can have the same energy level, and the spins of these two electrons must be in opposite directions. Different bonding characteristics for differ-ent elements exist because of the combined effects of electronic energy quantization and the limitation on the number of electrons at each energy level. An atom may be represented in simplified form by a small nucleus surrounded by diffuse concentric ‘‘clouds’’ of electrons (Fig. 3.3). The maximum num-ber of electrons that may be located in each diffuse shell is determined by quantum theory. The number and arrangement of electrons in the outermost shell are of prime importance for the development of different types of interatomic bonding and crystal structure. Interatomic bonds form when electrons in adjacent atoms interact in such a way that their energy levels are lowered. If the energy reduction is large, then a strong, primary bond develops. The way in which the bonding electrons are localized in space determines whether or not the bonds are directional. The strength and directionality of interatomic bonds, together with the relative sizes of the bonded atoms, determine the type of crystal structure assumed by a given compo-sition. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 5. SECONDARY BONDS 39 three limiting types: covalent, ionic, and metallic. They differ because of how the bonding electrons are local-ized in space. The energy of these bonds per mole of bonded atoms is from 60 103 to more than 400 103 joules (J; 15 to 100 kcal). As there are 6.023 1023 molecules per mole, it might be argued that such bonds are weak; however, relative to the weight of an atom they are very large. Covalent Bonds In the covalent bond, one or more bonding electrons are shared by two atomic nuclei to complete the outer shell for each atom. Covalent bonds are common in gases. If outer shell electrons are rep-resented Copyrighted Material by dots, then examples for (1) hydrogen gas, (2) methane, and (3) chlorine gas are: 1. H H H:H H 2. C 4H H:C:H H 3. :Cl Cl: :Cl:Cl: In the solid state, covalent bonds form primarily be-tween nonmetallic atoms such as oxygen, chlorine, nitrogen, and fluorine. Since only certain electrons participate in the bonding, covalent bonds are direc-tional. As a result, atoms bonded covalently pack in such a way that there are fixed bond angles. Ionic Bonds Ionic bonds form between positively and negatively charged free ions that acquire their charge through gain or loss of electrons. Cations (pos-itively charged atoms that are attracted by the cathode in an electric field) form by atoms giving up one or more loosely held electrons that lie outside a com-pleted electron shell and have a high energy level. Met-als, alkalies (e.g., sodium, potassium), and alkaline earths (e.g., calcium, magnesium) form cations. Anions (negatively charged atoms that are attracted to the an-ode) are those atoms requiring only a few electrons to complete their outer shell. Because the outer shells of ions are complete, structures cannot form by electron sharing as in the case of the covalent bond. Since ions are electrically charged, however, strong electrical at-tractions (and repulsions) can develop between them. The ionic bond is nondirectional. Each cation at-tracts all neighboring anions. In sodium chloride, which is one of the best examples of ionic bonding, a sodium cation attracts as many chlorine anions as will fit around it. Geometric considerations and electrical neutrality determine the actual arrangement of ioni-cally bonded atoms. As ionic bonding causes a separation between the centers of positive and negative charge in a molecule, the molecule will orient in an electrical field forming a dipole. The strength of this dipole is expressed in terms of the dipole moment . If two electrical charges of magnitude e, where e is the electronic charge, are separated by a distance d, then d e (3.1) Covalently bonded atoms may also produce dipolar molecules. Metallic Bonds Metals contain loosely held val-ence electrons that hold the positive metal ions to-gether but are free to travel through the solid material. Metallic bonds are nondirectional and can exist only among a large group of atoms. It is the large group of electrons and their freedom to move that make metals such good conductors of electricity and heat. The me-tallic bond is of little importance in most soils. Bonding in Soil Minerals A combination of ionic and covalent bonding is typical in most nonmetallic solids. Purely ionic or covalent bonding is a limiting condition that is the exception rather than the rule in most cases. Silicate minerals are the most abundant constituents of most soils. The in-teratomic bond in silica (SiO2) is about half covalent and half ionic. 3.4 SECONDARY BONDS Secondary bonds that are weak relative to ionic and covalent bonds also form between units of matter. They may be strong enough to determine the final arrange-ments of atoms in solids, and they may be sources of attraction between very small particles and between liquids and solid particles. The Hydrogen Bond If a hydrogen ion forms the positive end of a dipole, then its attraction to the negative end of an adjacent molecule is termed a hydrogen bond. Hydrogen bonds form only between strongly electronegative atoms such as oxygen and fluorine because these atoms produce the strongest dipoles. When the electron is detached from a hydrogen atom, such as when it combines with oxygen to form water, only a proton remains. As the electrons shared between the oxygen and hydrogen at-oms spend most of their time between the atoms, the oxygens act as the negative ends of dipoles, and the hydrogen protons act as the positive ends. The positive and negative ends of adjacent water molecules tie them together forming water and ice. The strength of the hydrogen bond is much greater than that of other secondary bonds because of the small size of the hydrogen ion. Hydrogen bonds are impor- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 6. 40 3 SOIL MINERALOGY within the lattice where atoms or atomic groups Copyrighted Material Figure 3.4 Examples of some common crystals. (hkl) are cleavage plane indices. From Dana’s Manual of Mineralogy, by C. S. Hurlbut, 16th Edition. Copyright 1957 by John Wiley Sons. Reprinted with permission from John Wiley Sons. tant in determining some of the characteristics of the clay minerals and in the interaction between soil par-ticle surfaces and water. van der Waals Bonds Permanent dipole bonds such as the hydrogen bond are directional. Fluctuating dipole bonds, commonly termed van der Waals bonds, also exist because at any one time there may be more electrons on one side of the atomic nucleus than on the other. This creates weak instantaneous dipoles whose oppositely charged ends attract each other. Although individual van der Waals bonds are weak, typically an order of magnitude weaker than a hydro-gen bond, they are nondirectional and additive between atoms. Consequently, they decrease less rapidly with distance than primary valence and hydrogen bonds when there are large groups of atoms. They are strong enough to determine the final arrangements of groups of atoms in some solids (e.g., many polymers), and they may be responsible for small cohesions in fine-grained soils. Van der Waals forces are described fur-ther in Chapter 7. 3.5 CRYSTALS AND THEIR PROPERTIES Particles composed of mineral crystals form the greatest proportion of the solid phase of a soil. A crys-tal is a homogeneous body bounded by smooth plane surfaces that are the external expression of an orderly internal atomic arrangement. A solid without internal atomic order is termed amorphous. Crystal Formation Crystals may form in three ways: 1. From Solution Ions combine as they separate from solution and gradually build up a solid of definite structure and shape. Halite (sodium chlo-ride) and other evaporites are examples. 2. By Fusion Crystals form directly from a liquid as a result of cooling. Examples are igneous rock minerals solidified from molten rock magma and ice from water. 3. From Vapor Although not of particular impor-tance in the formation of soil minerals, crystals can form directly from cooling vapors. Examples include snowflakes and flowers of sulfur. Examples of some common crystals are shown in Fig. 3.4. Characteristics of Crystals Certain crystal characteristics are used to distinguish different classes or groups of minerals. Variations in these characteristics result in different properties. 1. Structure The atoms in a crystal are arranged in a definite orderly manner to form a three-dimensional network termed a lattice. Positions Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 7. CRYSTALS AND THEIR PROPERTIES 41 are located are termed lattice points. Only 14 dif-ferent arrangements of lattice points in space are possible. These are the Bravais space lattices, and they are illustrated in Fig. 3.5. The smallest subdivision of a crystal that still possesses the characteristic composition and spa-tial same substance are constant. Crystals break along smooth cleavage planes. Cleavage planes lie between planes in which the atoms are most Copyrighted Material arrangement of atoms in the crystal is the unit cell. The unit cell is the basic repeating unit of the space lattice. 2. Cleavage and Outward Form The angles be-tween corresponding faces on crystals of the Figure 3.5 Unit cells of the 14 Bravais space lattices. The capital letters refer to the type of cell: P, primitive cell; C, cell with a lattice point in the center of two parallel faces; F, cell with a lattice point in the center of each face; I, cell with a lattice point in the center of the interior; R, rhombohedral primitive cell. All points indicated are lattice points. There is no general agreement on the unit cell to use for the hexagonal Bravais lattice; some prefer the P cell shown with solid lines, and others prefer the C cell shown in dashed lines (modified from Moffatt et al., 1965). Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 8. 42 3 SOIL MINERALOGY Copyrighted Material Figure 3.6 The six crystal systems. densely packed. This is because the center-to-center distance between atoms on opposite sides of the plane is greater than along other planes through the crystal. As a result, the strength along cleavage planes is less than in other directions. 3. Optical Properties The specific atomic arrange-ments within crystals allow light diffraction and polarization. These properties are useful for iden-tification and classification. Identification of rock minerals by optical means is common. Optical studies in soil are less useful because of the small sizes of most soil particles. 4. X-ray and Electron Diffraction The orderly atomic arrangements in crystals cause them to behave with respect to X-ray and electron beams in much the same way as does a diffraction grat-ing with respect to visible light. Different crystals yield different diffraction patterns. This makes X-ray diffraction a powerful tool for the study and identification of very small particles, such as clay that cannot be seen using optical means. 5. Symmetry There are 32 distinct crystal classes based on symmetry considerations involving the arrangement and orientation of crystal faces. These 32 classes may be grouped into 6 crystal systems with the classes within each system bear-ing close relationships to each other. The six crystal systems are illustrated in Fig. 3.6. Crystallographic axes parallel to the intersection edges of prominent crystal faces are established for each of the six crystal systems. In most crystals, these axes will also be symmetry axes or axes normal to symmetry planes. In five of the six systems, the crystals are re-ferred to three crystallographic axes. In the sixth (the hexagonal system), four axes are used. The axes are denoted by a, b, c (a1, a2, a3, and c in the hexagonal system) and the angles between the axes by , , and . Isometric or Cubic System There are three mutu-ally perpendicular axes of equal length. Mineral examples are galena, halite, magnetite, and pyrite. Hexagonal System Three equal horizontal axes ly-ing in the same plane intersect at 60 with a fourth axis perpendicular to the other three and of dif-ferent length. Examples are quartz, brucite, cal-cite, and beryl. Tetragonal System There are three mutually per-pendicular axes, with two horizontal of equal length, but different than that of the vertical axis. Zircon is an example. Orthorhombic System There are three mutually perpendicular axes, each of different length. Ex-amples include sulfur, anhydrite, barite, diaspore, and topaz. Monoclinic System There are three unequal axes, two inclined to each other at an oblique angle, with the third perpendicular to the other two. Ex-amples are orthoclase feldspar, gypsum, musco-vite, biotite, gibbsite, and chlorite. Triclinic System Three unequal axes intersect at oblique angles. Examples are plagioclase feldspar, kaolinite, albite, microcline, and turquoise. 3.6 CRYSTAL NOTATION Miller indices are used to describe plane orientations and directions in a crystal. This information, along with the distances that separate parallel planes is im-portant for the identification and classification of dif-ferent minerals. All lengths are expressed in terms of unit cell lengths. Any plane through a crystal may be described by intercepts, in terms of unit cell lengths, on the three or four crystallographic axes for the sys-tem in which the crystal falls. The reciprocals of these intercepts are used to index the plane. Reciprocals are used to avoid fractions and to account for planes par-allel to an axis (an intercept of infinity equals an index value of 0). Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 9. CRYSTAL NOTATION 43 lengths. Take plane mnp in Fig. 3.7a as an example. The intercepts of this plane are a 1, b 1, and c 1. The Miller indices of this plane are found by taking the reciprocals of these intercepts and clearing of fractions. Thus, Reciprocals are 1/1, 1/1, 1/1 Miller indices are (111) The indices are always enclosed within parentheses and indicated in the order abc without commas. Paren- Copyrighted Material An example illustrates the determination and mean-ing of Miller indices. Consider the mineral muscovite, a member of the monoclinic system. It has unit cell dimensions of a 0.52 nanometers (nm), b 0.90 nm, c 2.0 nm, and 95 30. Both the compo-sition and crystal structure of muscovite are similar to those of some of the important clay minerals. The muscovite unit cell dimensions and intercepts are shown in Fig. 3.7a. The intercepts for any plane of interest are first determined in terms of unit cell Figure 3.7 Miller indices: (a) Unit cell of muscovite, (b) (002) plane for muscovite, (c) (014) plane for muscovite, and (d) (623) plane for muscovite. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 10. 44 3 SOIL MINERALOGY Table 3.1 Atomic Packing, Structure, and Structural Stability Radius Ratioa Nb Geometry Example Stability 0–0.155 2 Line — — 0.155– 0.225 3 Triangle (CO3)2 Very high Material 0.225– 0.414 Copyrighted 4 Tetrahedron (SiO4)4 Moderately high 0.414– 0.732 6 Octahedron [Al(OH)6]3 High 0.732– 1.0 8 Body-cen-tered cube Iron Low 1.0 12 Sheet K–O bond in mica Very low aRange of cation to anion diameter ratios over which stable coordination is expected. bCoordination number. Table 3.2 Relative Stabilities of Some Soil Mineral Structural Units Structural Unit Approximate Relative Bond Strength (Valence/N) Silicon tetrahedron, (SiO4)4 4/4 1 Aluminum tetrahedron, [Al(OH)4]1 3/4 Aluminum octahedron, [Al(OH)6]3 3/6 1/2 Magnesium octahedron, [Mg(OH)6]4 2/6 1/3 K–O12 23 1/12 theses are always used to indicate crystallographic planes, whereas brackets are used to indicate direc-tions. For example, [111] designates line oq in Fig. 3.7a. Additional examples of Miller indices for planes through the muscovite crystal are shown in Figs. 3.7b, 3.7c, and 3.7d. A plane that cuts a negative axis is designated by placing a bar over the index that pertains to the negative intercept (Fig. 3.7d). The general index (hkl) is used to refer to any plane that cuts all three axes. Similarly (h00) designates a plane cutting only the a axis, (h0l) designates a plane parallel to the b axis, and so on. For crystals in the hexagonal system, the Miller index contains four numbers. The (001) planes of soil minerals are of particular interest be-cause they are indicative of specific clay mineral types. 3.7 FACTORS CONTROLLING CRYSTAL STRUCTURES Organized crystal structures do not develop by chance. The most stable arrangement of atoms in a crystal is that which minimizes the energy per unit volume. This is achieved by preserving electrical neutrality, satisfy-ing bond directionality, minimizing strong ion repul-sions, and packing atoms closely together. If the interatomic bonding is nondirectional, then the relative atomic sizes have a controlling influence on packing. The closest possible packing will maximize the number of bonds per unit volume and minimize the bonding energy. If interatomic bonds are directional, as is the case for covalent bonds, then both bond angles and atomic size are important. Anions are usually larger than cations because of electron transfer from cations to anions. The number of nearest neighbor anions that a cation possesses in a structure is termed the coordination number (N) or li-gancy. Possible values of coordination number in solid structures are 1 (trivial), 2, 3, 4, 6, 8, and 12. The relationships between atomic sizes, expressed as the ratio of cationic to anionic radii, coordination number, and the geometry formed by the anions are indicated in Table 3.1. Most solids do not have bonds that are completely nondirectional, and the second nearest neighbors may influence packing as well as the nearest neighbors. Even so, the predicted and observed coordination num-bers are in quite good agreement for many materials. The valence of the cation divided by the number of coordinated anions is an approximate indication of the relative bond strength, which, in turn, is related to the structural stability of the unit. Some of the structural units common in soil minerals and their relative bond strengths are listed in Table 3.2. The basic coordination polyhedra are seldom elec-trically neutral. In crystals formed by ionic bonded pol-yhedra, the packing maintains electrical neutrality and minimizes strong repulsions between ions with like charge. In such cases, the valence of the central cation equals the total charge of the coordinated anions, and the unit is really a molecule. Units of this type are held together by weaker, secondary bonds. An example is brucite, a mineral that has the composition Mg(OH)2. The Mg2 ions are in octahedral coordination with six (OH) ions forming a sheet structure in such a way that each (OH) is shared by 3Mg2. In a sheet con-taining N Mg2 ions, therefore, there must be 6N/3 2N (OH) ions. Thus, electrical neutrality results, and the sheet is in reality a large molecule. Successive oc- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 11. SURFACES 45 tahedral sheets are loosely bonded by van der Waals forces. Because of this, brucite has perfect basal cleav-age parallel to the sheets. Cations concentrate their charge in a smaller volume than do anions, so the repulsion between cations is greater than between anions. Cationic repulsions are minimized when the anions are located at the centers of coordination polyhedra. If the cations have a low valence, then the anion polyhedra pack as closely as possible to minimize energy per unit volume. If, on the other hand, the cations are small and highly charged, then the units arrange in a variety of ways in response to the repulsions. The silicon cation is in this category. Copyrighted Material 3.8 SILICATE CRYSTALS Small cations form structures with coordination num-bers of 3 and 4 (Table 3.1). These cations are often highly charged and generate strong repulsions between adjacent triangles or tetrahedra. As a result, such struc-tures share only corners and possibly edges, but never faces, since to do so would bring the cations too close together. The radius of silicon is only 0.039 nm, whereas that of oxygen is 0.132 nm. Thus silicon and oxygen combine in tetrahedral coordination, with the silicon occupying the space at the center of the tetra-hedron formed by the four oxygens. The tetrahedral arrangement satisfies both the directionality of the bonds (the Si–O bond is about half covalent and half ionic) and the geometry imposed by the radius ratio. Silicon is very abundant in Earth’s crust, amounting to about 25 percent by weight, but only 0.8 percent by volume. Almost half of igneous rock by weight and 91.8 percent by volume is oxygen. Silica tetrahedra join only at their corners, and sometimes not at all. Thus many crystal structures are possible, and there is a large number of silicate min-erals. Silicate minerals are classified according to how the silica tetrahedra (SiO4)4 associate with each other, as shown in Fig. 3.8. The tetrahedral combinations in-crease in complexity from the beginning to the end of the figure. The structural stability increases in the same direction. Island (independent) silicates are those in which the tetrahedra are not joined to each other. Instead, the four excess oxygen electrons are bonded to other positive ions in the crystal structure. In the olivine group, the minerals have the composition R2 2 SiO4. Garnets 4 contain cations of different valences and coordination numbers R2 R3(SiO). The negative charge of the 3 2 43SiO4 group in zircon is all balanced by the single Zr4. Ring and chain silicates are formed when corners of tetrahedra are shared. The formulas for these structures contain (SiO3)2. The pyroxene minerals are in this class. Enstatite, MgSiO3, is a simple member of this group. Some of the positions normally occupied by Si4 in single-chain structures may be filled by Al3. Substitution of ions of one kind by ions of another type, having either the same or different valence, but the same crystal structure, is termed isomorphous sub-stitution. The term substitution implies a replacement whereby a cation in the structure is replaced at some time by a cation of another type. In reality, however, the replaced cations were never there, and the mineral was formed with its present proportions of the different cations in the structure. Double chains of indefinite length may form with (Si4O11)6 as part of the structure. The amphiboles fall into this group (Fig. 3.8). Hornblendes have the same basic structure, but some of the Si4 positions are filled by Al3. The cations Na and K can be incorporated into the structure to satisfy electrical neutrality; Al3, Fe3, Fe2, and Mn2 can replace part of the Mg2 in sixfold coordination, and the (OH) group can be re-placed by F. In sheet silicates three of the four oxygens of each tetrahedron are shared to give structures containing (Si2O5)2. The micas, chlorites, and many of the clay minerals contain silica in a sheet structure. Framework silicates result when all four of the oxygens are shared with other tetrahedra. The most common example is quartz. In quartz, the silica tetrahedra are grouped to form spirals. The feldspars also have three-dimensional framework structures. Some of the silicon positions are filled by aluminum, and the excess negative charge thus created is balanced by cations of high coordina-tion such as potassium, calcium, sodium, and barium. Differences in the amounts of this isomorphous sub-stitution are responsible for the different members of the feldspar family. 3.9 SURFACES All liquids and solids terminate at a surface, or phase boundary, on the other side of which is matter of a different composition or state. In solids, atoms are bonded into a three-dimensional structure, and the ter-mination of this structure at a surface, or phase bound-ary, produces unsatisfied force fields. In a fine-grained particulate material such as clay soil the surface area may be very large relative to the mass of the material, and, as is emphasized throughout this book, the influ-ences of the surface forces on properties and behavior may be very large. Unsatisfied forces at solid surfaces may be balanced in any of the following ways: Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 12. 46 3 SOIL MINERALOGY Copyrighted Material Figure 3.8 Silica tetrahedral arrangements in different silicate mineral structures. Reprinted Gillott (1968) with permission from Elsevier Science Publishers BV. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 13. SURFACES 47 Material Copyrighted Figure 3.8 (Continued) 1. Attraction and adsorption of molecules from the adjacent phase 2. Cohesion with the surface of another mass of the same substance 3. Solid-state adjustments of the structure beneath the surface. Each unsatisfied bond force is significant relative to the weight of atoms and molecules. The actual mag-nitude of 1011 N or less, however, is infinitesimal compared to the weight of a piece of gravel or a grain of sand. On the other hand, consider the effect of re-ducing particle size. A cube 10 mm on an edge has a Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 14. 48 3 SOIL MINERALOGY surface area of 6.0 104 m2. If it is cut in half in the three directions, eight cubes result, each 5 mm on an edge. The surface area now is 12.0 104 m2. If the cubes are further divided to 1 m on an edge, the surface becomes 6.0 m2 for the same 1000 mm3 of material. Thus, as a solid is subdivided into smaller and smaller units, the proportion of surface area to weight becomes larger and larger. For a given particle shape, the ratio of surface area to volume is inversely proportional to some effective particle diameter. For many materials when particle size is reduced to 1 or 2 m or less the surface forces begin to exert a distinct influence on the behavior. Study of the behav-ior Copyrighted Material of particles of this size and less requires consider-ations of colloidal and surface chemistry. Most clay particles behave as colloids, both because of their small size and because they have unbalanced surface electrical forces as a result of isomorphous substitu-tions within their structure. Montmorillonite, which is one of the members of the smectite clay mineral group (see Section 3.17), may break down into particles that are only 1 unit cell thick (1.0 nm) when in a dispersed state and have a specific surface area of 800 m2 /g. If all particles con-tained in about 10 g of this clay could be spread out side by side, they would cover a football field. 3.10 GRAVEL, SAND, AND SILT PARTICLES The physical characteristics of cohesionless soils, that is, gravel, sand, and nonplastic silts, are determined primarily by particle size, shape, surface texture, and size distribution. The mineral composition determines hardness, cleavage, and resistance to physical and chemical breakdown. Some carbonate and sulfate min-erals, such as calcite and gypsum, are sufficiently sol-uble that their decomposition may be significant within the time frame of many projects. In many cases, how-ever, the nonclay particles may be treated as relatively inert, with interactions that are predominantly physical in nature. Evidence of this is provided by the soils on the Moon. Lunar soils have a silty, fine sand gradation; however, their compositions are totally different than those of terrestrial soils of the same gradation. The engineering properties of the two materials are sur-prisingly similar, however. The gravel, sand, and most of the silt fraction in a soil are composed of bulky, nonclay particles. As most soils are the products of the breakdown of preexisting rocks and soils, they are weathering products. Thus, the predominant mineral constituents of any soil are those that are one or more of the following: 1. Very abundant in the source material 2. Highly resistant to weathering, abrasion, and im-pact 3. Weathering products The nonclays are predominantly rock fragments or mineral grains of the common rock-forming minerals. In igneous rocks, which are the original source mate-rial for many soils, the most prevalent minerals are the feldspars (about 60 percent) and the pyroxenes and amphiboles (about 17 percent). Quartz accounts for about 12 percent of these rocks, micas for 4 percent, and other minerals for about 8 percent. However, in most soils, quartz is by far the most abundant mineral, with small amounts of feldspar and mica also present. Pyroxenes and amphiboles are sel-dom found in significant amounts. Carbonate minerals, mainly calcite and dolomite, are also found in some soils and can occur as bulky particles, shells, precipi-tates, or in solution. Carbonates dominate the compo-sition of some deep-sea sediments. Sulfates, in various forms, are found primarily in soils of semiarid and arid regions, with gypsum (CaSO4 2H2O) being the most common. Iron and aluminum oxides are abundant in residual soils of tropical regions. Quartz is composed of silica tetrahedra grouped to form spirals, with all tetrahedral oxygens bonded to silicon. The tetrahedral structure has a high stability. In addition, the spiral grouping of tetrahedra produces a structure without cleavage planes, quartz is already an oxide, there are no weakly bonded ions in the struc-ture, and the mineral has high hardness. Collectively, these factors account for the high persistence of quartz in soils. Feldspars are silicate minerals with a three-dimensional framework structure in which part of the silicon is replaced by aluminum. The excess negative charge resulting from this replacement is balanced by cations such as potassium, calcium, sodium, strontium, and barium. As these cations are relatively large, their coordination number is also large. This results in an open structure with low bond strengths between units. Consequently, there are cleavage planes, the hardness is only moderate, and feldspars are relatively easily broken down. This accounts for their lack of abun-dance in soils compared to their abundance in igneous rocks. Mica has a sheet structure composed of tetrahedral and octahedral units. Sheets are stacked one on the other and held together primarily by potassium ions in 12-fold coordination that provide an electrostatic bond of moderate strength. In comparison with the intralayer bonds, however, this bond is weak, which accounts for the perfect basal cleavage of mica. As a result of the Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 15. STRUCTURAL UNITS OF THE LAYER SILICATES 49 thin-plate morphology of mica flakes, sand and silts containing only a few percent mica may exhibit high compressibility when loaded and large swelling when unloaded, as may be seen in Fig. 3.9. The amphiboles, pyroxenes, and olivine have crystal structures that are rapidly broken down by weathering; hence they are absent from most soils. Some examples of silt and sand particles from dif-ferent Copyrighted Material soils are shown in Fig. 3.10. Angularity and roundness can be used to describe particle shapes, as shown in Fig. 3.11. Elongated and platy particles can develop preferred orientations, which can be respon-sible for anisotropic properties within a soil mass. The surface texture of the grains influences the stress– deformation and strength properties. 3.11 SOIL MINERALS AND MATERIALS FORMED BY BIOGENIC AND GEOCHEMICAL PROCESSES Evaporite deposits formed by precipitation of salts from salt lakes and seas as a result of the evaporation of water are sometimes found in layers that are several meters thick. The major constituents of seawater and their relative proportions are listed in Table 3.3. Also listed are some of the more important evaporite de-posits. Figure 3.9 Swelling index as a function of mica content for coarse-grained mixtures (data from Terzaghi, 1931). In some areas alternating layers of evaporite and clay or other fine-grained sediments are formed during cyclic wet and dry periods. Many limestones, as well as coral, have been formed by precipitation or from the remains of various organ-isms. Because of the much greater solubility of lime-stone than most other rock types, it may be the source of special problems caused by solution channels and cavities under foundations. Chemical sediments and rocks in freshwater lakes, ponds, swamps, and bays are occasionally encountered in civil engineering projects. Biochemical processes form marl, which ranges from relatively pure calcium carbonate to mixtures with mud and organic matter. Iron oxide is formed in some lakes. Diatomite or dia-tomaceous earth is essentially pure silica formed from the skeletal remains of small (up to a few tenths of a millimeter) freshwater and saltwater organisms. Owing to their solubility limestone, calcite, gypsum, and other salts may cause special geotechnical problems. Oxidation and reduction of pyrite-bearing earth ma-terials, that is, soils and rocks containing FeS2, can be the source of many types of geotechnical problems, including ground heave, high swell pressures, forma-tion of acid drainage, damage to concrete, and corro-sion of steel (Bryant et al., 2003). The chemical and biological processes and consequences of pyritic re-actions are covered in Sections 8.3, 8.11, and 8.16. More than 12 percent of Canada is covered by a peaty material, termed muskeg, composed almost en-tirely of decaying vegetation. Peat and muskeg may have water contents of 1000 percent or more; they are very compressible, and they have low strength. The special properties of these materials and methods for analysis of geotechnical problems associated with them are given by MacFarlane (1969), Dhowian and Edil (1980), and Edil and Mochtar (1984). 3.12 SUMMARY OF NONCLAY MINERAL CHARACTERISTICS Important compositional, structural, and morphological characteristics of the important nonclay minerals found in soils are summarized in Table 3.4. Of these miner-als, quartz is by far the most common, both in terms of the number of soils in which it is found and its abundance in a typical soil. Feldspar and mica are fre-quently present in small percentages. 3.13 STRUCTURAL UNITS OF THE LAYER SILICATES Clay minerals in soils belong to the mineral family termed phyllosilicates, which also contains other layer Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 16. 50 3 SOIL MINERALOGY Material Copyrighted Figure 3.10 Photomicrographs of sand and silt particles from several soils: (a) Ottawa stan-dard sand, (b) Monterey sand, (c) Sacramento River sand, (d) Eliot sand, and (e) lunar soil mineral grains (photo courtesy Johnson Space Center). Squares in background area are 11 mm. (ƒ) Recrystallized breccia particles from lunar soil (photo courtesy of NASA Johnson Space Center). Squares in background grid are 11 mm. silicates such as serpentine, pyrophyllite, talc, mica, and chlorite. Clay minerals occur in small particle sizes, and their unit cells ordinarily have a residual negative charge that is balanced by the adsorption of cations from solution. The structures of the common layer silicates are made up of combinations of two simple structural units, the silicon tetrahedron (Fig. 3.12) and the alu-minum or magnesium octahedron (Fig. 3.13). Different clay mineral groups are characterized by the stacking arrangements of sheets1 (sometimes chains) of these 1 In conformity with the nomenclature of the Clay Minerals Society (Bailey et al., 1971), the following terms are used: a plane of atoms, a sheet of basic structural units, and a layer of unit cells composed of two, three, or four sheets. units and the manner in which two successive two- or three-sheet layers are held together. Differences among minerals within clay mineral groups result primarily from differences in the type and amount of isomorphous substitution within the crystal structure. Possible substitutions are nearly endless in number, and the crystal structure arrangement may range from very poor to nearly perfect. Fortunately for engineering purposes, knowledge of the structural and compositional characteristics of each group, without detailed study of the subtleties of each specific mineral, is adequate. Silica Sheet In most clay mineral structures, the silica tetrahedra are interconnected in a sheet structure. Three of the Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 17. STRUCTURAL UNITS OF THE LAYER SILICATES 51 Material Copyrighted Figure 3.10 (Continued) Figure 3.11 Sand and silt size particle shapes as seen in silhouette. four oxygens in each tetrahedron are shared to form a hexagonal net, as shown in Figs. 3.12b and 3.14. The bases of the tetrahedra are all in the same plane, and the tips all point in the same direction. The structure has the composition (Si4O10)4 and can repeat indefi-nitely. Electrical neutrality can be obtained by replace-ment of four oxygens by hydroxyls or by union with a sheet of different composition that is positively charged. The oxygen-to-oxygen distance is 2.55 ang-stroms (A˚ ),2 the space available for the silicon ion is 0.55 A˚ , and the thickness of the sheet in clay mineral structures is 4.63 A˚ (Grim, 1968). 2 In conformity with the SI system of units, lengths should be given in nanometers. For convenience, however, the angstrom unit is re-tained for atomic dimensions, where 1 A˚ 0.1 nm. Silica Chains In some of the less common clay minerals, silica tet-rahedra are arranged in bands made of double chains of composition (Si4O11)6. Electrical neutrality is achieved and the bands are bound together by alumi-num and/or magnesium ions. A diagrammatic sketch of this structure is shown in Fig. 3.8. Minerals in this group resemble the amphiboles in structure. Octahedral Sheet This sheet structure is composed of magnesium or alu-minum in octahedral coordination with oxygens or hy-droxyls. In some cases, other cations are present in place of Al3 and Mg2, such as Fe2, Fe3, Mn2, Ti4, Ni2, Cr3, and Li. Figure 3.13b is a schematic diagram of such a sheet structure. The oxygen-to-oxygen distance is 2.60 A˚ , and the space available for the octahedrally coordinated cation is 0.61 A˚ . The thickness of the sheet is 5.05 A˚ in clays (Grim, 1968). If the cation is trivalent, then normally only two-thirds of the possible cationic spaces are filled, and the structure is termed dioctahedral. In the case of alu-minum, the composition is Al2(OH)6. This composition and structure form the mineral gibbsite. When com-bined with silica sheets, as is the case in clay mineral Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 18. 52 3 SOIL MINERALOGY Table 3.3 Major Constituents of Seawater and Evaporite Deposits Ion Grams per Liter Percent by Weight of Total Solids Important Evaporite Deposits Sodium, Na 10.56 30.61 Anhydrite CaSO4 Magnesium, Mg2 1.27 3.69 Barite BaSO4 Calcium, Ca2 0.40 1.16 Celesite SrSO4 Potassium, K 0.38 1.10 Kieserite MgSO4 H2O Strontium, Sr2 0.013 0.04 Gypsum CaSO4 2H2O Chloride, Cl 18.98 55.04 Polyhalite Ca2K2Mg(SO4) 2H2O Sulfate, SO4 Copyrighted Material 2 2.65 7.68 Bloedite Ma2Mg(SO4)2 4H2O Bicarbonate, HCO3 0.14 0.41 Hexahydrite MgSO4 6H2O Bromide, Br 0.065 0.19 Epsomite MgSO4 7H2O Fluoride, F 0.001 — Kainite K4Mg4(Cl/SO4) 1 1H2 O Boric Acid, H3BO3 0.026 0.08 Halite NaCl 34.485 100.00 Sylvite KCl Flourite CaF2 Bischofite MgCl2 6H2O Carnallite KMgCl3 6H2O Adapted from data by Degens (1965). structures, an aluminum octahedral sheet is referred to as a gibbsite sheet. If the octahedrally coordinated cation is divalent, then normally all possible cation sites are occupied and the structure is trioctahedral. In the case of magne-sium, the composition is Mg3(OH)6, giving the mineral brucite. In clay mineral structures, a sheet of magne-sium octahedra is termed a brucite sheet. Schematic representations of the sheets are useful for simplified diagrams of the structures of the differ-ent clay minerals: Silica sheet or Octahedral sheet (Various cations in octahedral coordination) Gibbsite sheet (Octahedral sheet cations are mainly aluminum) Brucite sheet (Octahedral sheet cations are mainly magnesium) Water layers are found in some structures and may be represented by for each molecular layer. Atoms of a specific type, for example, potassium, are represented thus: K . The diagrams are indicative of the clay mineral layer structure. They do not indicate the correct width-to-length ratios for the actual particles. The structures shown are idealized; in actual minerals, irregular sub-stitutions and interlayering or mixed-layer structures are common. Furthermore, direct assembly of the basic units does not necessarily form the naturally occurring minerals. The ‘‘building block’’ approach is useful, however, for the development of conceptual models. 3.14 SYNTHESIS PATTERN AND CLASSIFICATION OF THE CLAY MINERALS The manner in which atoms are assembled into tetra-hedral and octahedral units, followed by the formation of sheets and their stacking to form layers that combine to produce the different clay mineral groups is illus-trated in Fig. 3.15. The basic structures shown in the bottom row of Fig. 3.15 comprise the great prepon-derance of the clay mineral types that are found in soils. Grouping the clay minerals according to crystal structure and stacking sequence of the layers is con-venient since members of the same group have gen- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 19. SYNTHESIS PATTERN AND CLASSIFICATION OF THE CLAY MINERALS 53 Table 3.4 Properties and Characteristics of Nonclay Minerals in Soils Mineral Formula Crystal System Cleavage Particle Shape Specific Gravity Hardness Occurrence in Soils of Engineering Interest Quartz SiO2 Hexagonal None Bulky 2.65 7 Very abundant Orthoclase feldspar Copyrighted Material KalSi3O8 Monoclinic 2 planes Elongate 2.57 6 Common Plagioclase feldspar NaAlSi3O8 CaAl2Si3O8 (variable) Triclinic 2 planes Bulky— elongate 2.62–2.76 6 Common Muscovite mica Kal3Si3O10(OH)2 Monoclinic Perfect basal Thin plates 2.76–3.1 2–21⁄2 Common Biotite mica K(Mg,FE)3AlSi3O10(OH)2 Monoclinic Perfect basal Thin plates 2.8–3.2 21⁄2–3 Common Hornblende Na,Ca,Mg,Fe,Al silicate Monoclinic Perfect prismatic Prismatic 3.2 5–6 Uncommon Augite (pyroxene) Ca(Mg,Fe,Al)(Al,Si)2O6 Monoclinic Good prismatic Prismatic 3.2–3.4 5–6 Uncommon Olivine (Mg,Fe)2SiO4 Orthorhombic Conchoidal fracture Bulky 3.27–3.37 61⁄2–7 Uncommon Calcite CaCO3 Hexagonal Perfect Bulky 2.72 21⁄2–3 May be abundant locally Dolomite CaMg(CO3)2 Hexagonal Perfect rhombohedral Bulky 2.85 31⁄2–4 May be abundant locally Gypsum CaSO4 2H2O Monoclinic 4 planes Elongate 2.32 2 May be abundant locally Pyrite FeS2 Isometric Cubical Bulky cubic 5.02 6–61⁄2 Data from Hurlbut (1957). Figure 3.12 Silicon tetrahedron and silica tetrahedra arranged in a hexagonal network. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 20. 54 3 SOIL MINERALOGY Figure 3.13 Octahedral unit and sheet Material structure of octahedral units. Copyrighted Figure 3.14 Silica sheet in plan view. erally similar engineering properties. The minerals have unit cells consisting of two, three, or four sheets. The two-sheet minerals are made up of a silica sheet and an octahedral sheet. The unit layer of the three-sheet minerals is composed of either a dioctahedral or trioctahedral sheet sandwiched between two silica sheets. Unit layers may be stacked closely together or water layers may intervene. The four-sheet structure of chlorite is composed of a 21 layer plus an interlayer hydroxide sheet. In some soils, inorganic, claylike ma-terial is found that has no clearly identifiable crystal structure. Such material is referred to as allophane or noncrystalline clay. The bottom row of Fig. 3.15 shows that the 21 minerals differ from each other mainly in the type and amount of ‘‘glue’’ that holds the successive layers to-gether. For example, smectite has loosely held cations between the layers, illite contains firmly fixed potas-sium ions, and vermiculite has somewhat organized layers of water and cations. The chlorite group repre-sents an end member that has 21 layers bonded by an organized hydroxide sheet. The charge per formula Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 21. INTERSHEET AND INTERLAYER BONDING IN THE CLAY MINERALS 55 Material Figure 3.15 Synthesis pattern for the clay minerals. Copyrighted unit is variable both within and among groups, and reflects the fact that the range of compositions is great owing to varying amounts of isomorphous substitution. Accordingly, the boundaries between groups are some-what arbitrary. Isomorphous Substitution The concept of isomorphous substitution was intro-duced in Section 3.13 in connection with some of the silicate crystals. It is very important in the structure and properties of the clay minerals. In an ideal gibbsite sheet, only two-thirds of the octahedral positions are filled, and all of the cations are aluminum. In an ideal brucite sheet, all the octahedral spaces are filled by magnesium. In an ideal silica sheet, silicons occupy all tetrahedral spaces. In clay minerals, however, some of the tetrahedral and octahedral spaces are occupied by cations other than those in the ideal structure. Common examples are aluminum in place of silicon, magnesium instead of aluminum, and ferrous iron (Fe2) for mag-nesium. This presence in an octahedral or tetrahedral position of a cation other than that normally found, without change in crystal structure, is isomorphous substitution. The actual tetrahedral and octahedral cat-ion distributions may develop during initial formation or subsequent alteration of the mineral. 3.15 INTERSHEET AND INTERLAYER BONDING IN THE CLAY MINERALS A single plane of atoms that are common to both the tetrahedral and octahedral sheets forms a part of the clay mineral layers. Bonding between these sheets is of the primary valence type and is very strong. How-ever, the bonds holding the unit layers together may be of several types, and they may be sufficiently weak that the physical and chemical behavior of the clay is influenced by the response of these bonds to changes in environmental conditions. Isomorphous substitution in all of the clay minerals, with the possible exception of those in the kaolinite group, gives clay particles a net negative charge. To preserve electrical neutrality, cations are attracted and held between the layers and on the surfaces and edges of the particles. Many of these cations are exchange-able cations because they may be replaced by cations of another type. The quantity of exchangeable cations is termed the cation exchange capacity (cec) and is usually expressed as milliequivalents (meq)3 per 100 g of dry clay. Five types of interlayer bonding are possible in the layer silicates (Marshall, 1964). 1. Neutral parallel layers are held by van der Waals forces. Bonding is weak; however, stable crystals of appreciable thickness such as the nonclay min- 3Equivalent weight combining weight of an element (atomic weight /valence). Number of equivalents (weight of element / atomic weight) valence. The number of ions in an equivalent Avogardro’s number/valence. Avogadro’s number 6.02 1023. An equivalent contains 6.02 1023 electron charges or 96,500 coulombs, which is 1 faraday. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 22. 56 3 SOIL MINERALOGY erals of pyrophyllite and talc may form. These minerals cleave parallel to the layers. 2. In some minerals (e.g., kaolinite, brucite, gibb-site), there are opposing layers of oxygens and hydroxyls or hydroxyls and hydroxyls. Hydrogen bonding then develops between the layers as well as van der Waals bonding. Hydrogen bonds re-main stable in the presence of water. Copyrighted Material 3. Neutral silicate layers that are separated by highly polar water molecules may be held to-gether by hydrogen bonds. 4. Cations needed for electrical neutrality may be in positions that control interlayer bonding. In mi-cas, some of the silicon is replaced by aluminum in the silica sheets. The resulting charge defi-ciency is partly balanced by potassium ions be-tween the unit cell layers. The potassium ion just fits into the holes formed by the bases of the silica tetrahedra (Fig. 3.12). As a result, it gen-erates a strong bond between the layers. In the chlorites, the charge deficiencies from substitu-tions in the octahedral sheet of the 21 sandwich are balanced by excess charge on the single-sheet layer interleaved between the three-sheet layers. This provides a strongly bonded structure that while exhibiting cleavage will not separate in the presence of water or other polar liquids. 5. When the surface charge density is moderate, as in smectite and vermiculite, the silicate layers readily adsorb polar molecules, and also the ad-sorbed cations may hydrate, resulting in layer separation and expansion. The strength of the in-terlayer bond is low and is a strong function of charge distribution, ion hydration energy, surface ion configuration, and structure of the polar mol-ecule. Smectite and vermiculite particles adsorb water be-tween the unit layers and swell, whereas particles of the nonclay minerals, pyrophyllite and talc, which have comparable structures, do not. There are two possible reasons (van Olphen, 1977): 1. The interlayer cations in smectite hydrate, and the hydration energy overcomes the attractive forces between the unit layers. There are no in-terlayer cations in pyrophyllite; hence, no swell-ing. 2. Water does not hydrate the cations but is ad-sorbed on oxygen surfaces by hydrogen bonds. There is no swelling in pyrophyllite and talc be-cause the surface hydration energy is too small to overcome the van der Waals forces between layers, which are greater in these minerals be-cause of a smaller interlayer distance. Whatever the reason, the smectite minerals are the dominant source of swelling in the expansive soils that are so prevalent throughout the world. 3.16 THE 11 MINERALS The kaolinite–serpentine minerals are composed of al-ternating silica and octahedral sheets as shown sche-matically in Fig. 3.16. The tips of the silica tetrahedra and one of the planes of atoms in the octahedral sheet are common. The tips of the tetrahedra all point in the same direction, toward the center of the unit layer. In the plane of atoms common to both sheets, two-thirds of the atoms are oxygens and are shared by both sili-con and the octahedral cations. The remaining atoms in this plane are (OH) located so that each is directly below the hole in the hexagonal net formed by the bases of the silica tetrahedra. If the octahedral layer is brucite, then a mineral of the serpentine subgroup re-sults, whereas dioctahedral gibbsite layers give clay minerals in the kaolinite subgroup. Trioctahedral 11 minerals are relatively rare, usually occur mixed with kaolinite or illite, and are hard to identify. A diagram-matic sketch of the kaolinite structure is shown in Fig. 3.17. The structural formula is (OH)8Si4Al4O10, and the charge distribution is indicated in Fig. 3.18. Mineral particles of the kaolinite subgroup consist of the basic units stacked in the c direction. The bond-ing between successive layers is by both van der Waals forces and hydrogen bonds. The bonding is sufficiently strong that there is no interlayer swelling in the pres-ence of water. Because of slight differences in the oxygen-to-oxygen distances in the tetrahedral and octahedral lay-ers, there is some distortion of the ideal tetrahedral network. As a result, kaolinite, which is the most abun-dant member of the subgroup and a common soil min-eral, is triclinic instead of monoclinic. The unit cell dimensions are a 5.16 A˚ , b 8.94 A˚ , c 7.37 A˚ , 91.8, 104.5, and 90. Variations in stacking of layers above each other, and possibly in the position of aluminum ions within the available sites in the octahedral sheet, produce dif-ferent members of the kaolinite subgroup. The dickite unit cell is made up of two unit layers, and the nacrite unit cell contains six. Both appear to be formed by hydrothermal processes. Dickite is fairly common as secondary clay in the pores of sandstone and in coal beds. Neither dickite nor nacrite is common in soils. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 23. THE 11 MINERALS 57 Figure 3.16 Schematic diagrams of the structures ˚ Aof kaolinite and serpentine: (a) kaolinite and (b) serpentine. Material Copyrighted is about 10.1 . The difference between these Figure 3.17 Diagrammatic sketch of the kaolinite structure. Halloysite Halloysite is a particularly interesting mineral of the kaolinite subgroup. Two distinct endpoint forms of this mineral exist, as shown in Fig. 3.19; one, a hydrated form consisting of unit kaolinite layers separated from each other by a single layer of water molecules and having the composition (OH)8Si4Al4O10 4H2O, and the other, a nonhydrated form having the same unit layer structure and chemical composition as kaolinite. The basal spacing in the c direction d(001) for the non-hydrated form is about 7.2 A˚ , as for kaolinite. Because of the interleaved water layer, d(001) for hydrated hal-loysite Figure 3.18 Charge distribution on kaolinite. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 24. 58 3 SOIL MINERALOGY Figure 3.19 Schematic diagrams of the structure of halloysite: (a) halloysite (10 ) and (b) halloysite (7 ). Material Copyrighted varieties. ˚ A˚ AFigure 3.20 Electron photomicrograph of well-crystallized kaolinite from St. Austell, Cornwall, England. Picture width is 17 m (Tovey, 1971). values, 2.9 A˚ , is the approximate thickness of a single layer of water molecules. The recommended terms for the two forms of hal-loysite are halloysite (7 A˚ ) and halloysite (10 A˚ ). Transformation from halloysite (10 A˚ ) to halloysite (7 A˚ ) by dehydration can occur at relatively low temper-atures and is irreversible. Halloysite is often found in soils formed from volcanic parent materials in wet en-vironments. It can be responsible for special properties and problems in earthwork construction, as discussed later in this book. Isomorphous Substitution and Exchange Capacity Whether or not measurable isomorphous substitution exists within the structure of the kaolinite minerals is uncertain. Nevertheless, values of cation exchange ca-pacity in the range of 3 to 15 meq/100 g for kaolinite and from 5 to 40 meq/100 g for halloysite have been measured. Thus, kaolinite particles possess a net neg-ative charge. Possible sources are: 1. Substitution of Al3 for Si4 in the silica sheet or a divalent ion for Al3 in the octahedral sheet. Replacement of only 1 Si in every 400 would be adequate to account for the exchange capacity. 2. The hydrogen of exposed hydroxyls may be re-placed by exchangeable cations. According to Grim (1968), however, this mechanism is not likely because the hydrogen would probably not be replaceable under the conditions of most exchange reactions. 3. Broken bonds around particle edges may give un-satisfied charges that are balanced by adsorbed cations. Kaolinite particles are charged positively on their edges when in a low pH (acid) environment, but neg-atively charged in a high pH (basic) environment. Low exchange capacities are measured under low pH con-ditions and high exchange capacities are obtained for determinations at high pH. This suggests that broken bonds are at least a partial source of exchange capacity. That a positive cation exchange capacity is measured under low pH conditions when edges are positively charged indicates that some isomorphous substitution must exist also. As interlayer separation does not occur in kaolinite, balancing cations must adsorb on the exterior surfaces and edges of the particles. Morphology and Surface Area Well-crystallized particles of kaolinite (Fig. 3.20), na-crite, and dickite occur as well-formed six-sided plates. The lateral dimensions of these plates range from about 0.1 to 4 m, and their thicknesses are from about 0.05 to 2 m. Poorly crystallized kaolinite generally occurs as less distinct hexagonal plates, and the parti-cle size is usually smaller than for the well-crystallized Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 25. SMECTITE MINERALS 59 Halloysite (10 A˚ ) occurs as cylindrical tubes of overlapping sheets of the kaolinite type (Fig. 3.21). The c axis at any point nearly coincides with the tube radius. The formation of tubes has been attributed to a misfit in the b direction of the silica and gibbsite sheets (Bates et al., 1950). The b dimension in kaolinite is 8.93 A˚ ; in gibbsite it is only 8.62 A˚ . This means that the (OH) spacing in gibbsite sheets is stretched in order to obtain a fit with the silica sheet. Evidently, in hal-loysite Copyrighted Material (10 A˚ ), the reduced interlayer bond, caused by the intervening layer of water molecules, enables the (OH) layer to revert to 8.62 A˚ , resulting in a curvature with the hydroxyls on the inside and the bases of the silica tetrahedra on the outside. The outside diameters of the tubular particles range from about 0.05 to 0.20 m, with a median value of 0.07 m. The wall thick-ness is about 0.02 m. The tubes range in length from a fraction of a micrometer to several micrometers. Dry-ing of halloysite (10 A˚ ) may result in splitting or un-rolling of the tubes. The specific surface area of kaolinite is about 10 to 20 m2 /g of dry clay; that of halloysite (10 A˚ ) is 35 to 70 m2/g. 3.17 SMECTITE MINERALS Structure The minerals of the smectite group have a prototype structure similar to that of pyrophyllite, consisting of an octahedral sheet sandwiched between two silica sheets, as shown schematically in Fig. 3.22 and dia-grammatically in three dimensions in Fig. 3.23. All the tips of the tetrahedra point toward the center of the unit cell. The oxygens forming the tips of the tetra-hedra are common to the octahedral sheet as well. The anions in the octahedral sheet that fall directly above Figure 3.21 Electron photomicrograph of halloysite from Bedford, Indiana. Picture width is 2 m (Tovey, 1971). and below the hexagonal holes formed by the bases of the silica tetrahedra are hydroxyls. The layers formed in this way are continuous in the a and b directions and stacked one above the other in the c direction. Bonding between successive layers is by van der Waals forces and by cations that balance charge deficiencies in the structure. These bonds are weak and easily separated by cleavage or adsorption of water or other polar liquids. The basal spacing in the c direction, d(001), is variable, ranging from about 9.6 A˚ to complete separation. The theoretical composition in the absence of isomorphous substitutions is (OH)4Si8Al4O20 n(interlayer)H2O. The structural configuration and cor-responding charge distribution are shown in Fig. 3.24. The structure shown is electrically neutral, and the atomic configuration is essentially the same as that in the nonclay mineral pyrophyllite. Isomorphous Substitution in the Smectite Minerals Smectite minerals differ from pyrophyllite in that there is extensive isomorphous substitution for silicon and aluminum by other cations. Aluminum in the octahe-dral sheet may be replaced by magnesium, iron, zinc, nickel, lithium, or other cations. Aluminum may re-place up to 15 percent of the silicon ions in the tetra-hedral sheet. Possibly some of the silicon positions can be occupied by phosphorous (Grim, 1968). Substitutions for aluminum in the octahedral sheet may be one-for-one or three-for-two (aluminum oc-cupies only two-thirds of the available octahedral sites) in any combination from a few to complete replace-ment. The resulting structure, however, is either almost exactly dioctahedral (montmorillonite subgroup) or trioctahedral (saponite subgroup). The charge defi-ciency resulting from these substitutions ranges from 0.5 to 1.2 per unit cell. Usually, it is close to 0.66 per unit cell. A charge deficiency of this amount would result from replacement of every sixth aluminum by a magnesium ion. Montmorillonite, the most common mineral of the group, has this composition. Charge de-ficiencies that result from isomorphous substitution are balanced by exchangeable cations located between the unit cell layers and on the surfaces of particles. Some minerals of the smectite group and their com-positions are listed in Table 3.5. An arrow indicates the source of the charge deficiency, which has been as-sumed to be 0.66 per unit cell in each case. Sodium is indicated as the balancing cation. The formulas should be considered indicative of the general character of the mineral, but not as absolute, because a variety of com-positions can exist within the same basic crystal struc-ture. Because of the large amount of unbalanced Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 26. 60 3 SOIL MINERALOGY Figure 3.22 Schematic diagrams of the structures of the smectite minerals: (a) montmoril-lonite and (b) saponite. Material Copyrighted Figure 3.23 Diagrammatic sketch of the montmorillonite structure. Figure 3.24 Charge distribution in pyrophyllite (type struc-ture for montmorillonite). substitution in the smectite minerals, they have high cation exchange capacities, generally in the range of 80 to 150 meq/100 g. Morphology and Surface Area Montmorillonite may occur as equidimensional flakes that are so thin as to appear more like films, as shown in Fig. 3.25. Particles range in thickness from 1-nm unit layers upward to about 1/100 of the width. The long axis of the particle is usually less than 1 or 2 m. When there is a large amount of substitution of iron and/or magnesium for aluminum, the particles may be lath or needle shaped because the larger Mg2 and Fe3 ions cause a directional strain in the structure. The specific surface area of smectite can be very large. The primary surface area, that is, the surface area exclusive of interlayer zones, ranges from 50 to 120 m2 /g. The secondary specific surface that is exposed Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 27. SMECTITE MINERALS 61 Table 3.5 Some Minerals of the Smectite Group Mineral Tetrahedral Sheet Substitutions Octahedral Sheet Substitutions Formula/Unit Cella Dioctahedral, Smectites or Montmorillonites Montmorillonite None 1Mg2 for every sixth Al3 (OH)4Si8(Al3.34Mg0.66) O20 ↓ Na0.66 Copyrighted Material Beidellite Al for Si None (OH)4(Si6.34Al1.66) Al4.34O20 ↓ Na0.66 Nontronite Al for Si Fe3 for Al (OH)4(Si7.34Al0.66) Fe4 3O20 ↓ Na0.66 Trioctahedral, Smectites, or Saponites Hectorite None Li for Mg (OH)4Si8(Mg5.34Li0.66) O20 ↓ Na0.66 Saponite Al for Si Fe3 for Mg (OH)4(Si7.34Al0.66) Mg6O20 ↓ Na0.66 Sauconite Al for Si Zn for Mg (OH)4(Si8yAly)(Zn6xMgx) O20 ↓ Na0.66 aTwo formula units are needed to give one unit cell. After Ross and Hendricks (1945); Marshall (1964); and Warshaw and Roy (1961). Figure 3.25 Electron photomicrograph of montmorillonite (bentonite) from Clay Spur, Wyoming. Picture width is 7.5 m (Tovey, 1971). by expanding the lattice so that polar molecules can penetrate between layers can be up to 840 m2/g. Bentonite A very highly plastic, swelling clay material known as bentonite is very widely used for a variety of purposes, ranging from drilling mud and slurry walls to clarifi-cation of beer and wine. The bentonite familiar to most geoengineers is a highly colloidal, expansive alteration product of volcanic ash. It has a liquid limit of 500 percent or more. It is widely used as a backfill during the construction of slurry trench walls, as a soil ad-mixture for construction of seepage barriers, as a grout material, as a sealant for piezometer installations, and for other special applications. When present as a major constituent in soft shale or as a seam in rock formations, bentonite may be a cause of continuing slope stability problems. Slide problems at Portuguese Bend along the Pacific Ocean in southern California, in the Bearpaw shale in Saskatchewan, and in the Pierre shale in South Dakota are in large mea- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 28. 62 3 SOIL MINERALOGY sure due to the high content of bentonite. Stability problems in underground construction may be caused by the presence of montmorillonite in joints and faults (Brekke and Selmer-Olsen, 1965). 3.18 MICALIKE CLAY MINERALS Illite is the most commonly found clay mineral in soils encountered in engineering practice. Its structure is quite similar to that of muscovite mica, and it is some-times 3.26b. The actual thickness of the water layer depends on the cations that balance the charge deficiencies in Copyrighted Material referred to as hydrous mica. Vermiculite is also often found as a clay phase constituent of soils. Its structure is related to that of biotite mica. Structure The basic structural unit for the muscovite (white mica) is shown schematically in Fig. 3.26a. It is the three-layer silica–gibbsite–silica sandwich that forms pyro-phyllite, with the tips of all the tetrahedra pointing toward the center and common with octahedral sheet ions. Muscovite differs from pyrophyllite, however, in that about one-fourth of the silicon positions are filled by aluminum, and the resulting charge deficiency is balanced by potassium between the layers. The layers are continuous in the a and b directions and stacked in the c direction. The radius of the potassium ion, 1.33 , is such that it fits snugly in the 1.32 A˚ radius hole formed by the bases of the silica tetrahedra. It is in 12- fold coordination with the 6 oxygens in each layer. A diagrammatic three-dimensional sketch of the muscovite structure is shown in Fig. 3.27. The struc-tural Figure 3.26 Schematic diagram of the structures of muscovite, illite, and vermiculite: (a) muscovite and illite and (b) vermiculite. A˚ configuration and charge distribution are shown in Fig. 3.28. The unit cell is electrically neutral and has the formula (OH)4K2(Si6Al2)Al4O20. Muscovite is the dioctahedral end member of the micas and contains only Al3 in the octahedral layer. Phlogopite (brown mica) is the trioctahedral end member, with the octa-hedral positions filled entirely by magnesium. It has the formula (OH)4K2(Si6Al2)Mg6O20. Biotite (black mica) is trioctahedral, with the octahedral positions filled mostly by magnesium and iron. It has the general formula (OH)4K2(Si6Al2)(MgFe)6O20. The relative pro-portions of magnesium and iron may vary widely. Illite differs from mica in the following ways (Grim, 1968): 1. Fewer of the Si4 positions are filled by Al3 in illite. 2. There is some randomness in the stacking of lay-ers in illite. 3. There is less potassium in illite. Well-organized illite contains 9 to 10 percent K2O (Weaver and Pollard, 1973). 4. Illite particles are much smaller than mica parti-cles. Some illite may contain magnesium and iron in the octahedral sheet as well as aluminum (Marshall, 1964). Iron-rich illite, usually occurring as earthy green pel-lets, is termed glauconite. The vermiculite structure consists of regular inter-stratification of biotite mica layers and double molec-ular layers of water, as shown schematically in Fig. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 29. MICALIKE CLAY MINERALS 63 Material Copyrighted Figure 3.27 Diagrammatic sketch of the structure of muscovite. Figure 3.28 Charge distribution in muscovite. the biotitelike layers. With magnesium or calcium present, which is the usual case in nature, there are two water layers, giving a basal spacing of 14 A˚ . A general formula for vermiculite is (OH)4(MgCa)x(Si8 xAlx)(MgFe)6O20yH2O x 1 to 1.4 y 8 Isomorphous Substitution and Exchange Capacity There is extensive isomorphous substitution in illite and vermiculite. The charge deficiency in illite is 1.3 to 1.5 per unit cell. It is located primarily in the silica sheets and is balanced partly by the nonexchangeable potassium between layers. Thus, the cation exchange capacity of illite is less than that of smectite, amount-ing to 10 to 40 meq/100 g. Values greater than 10 to 15 meq/100 g may be indicative of some expanding layers (Weaver and Pollard, 1973). In the absence of fixed potassium the exchange capacity would be about 150 meq/100 g. Interlayer bonding by potassium is so strong that the basal spacing of illite remains fixed at 10 A˚ in the presence of polar liquids. The charge deficiency in vermiculite is 1 to 1.4 per unit cell. Since the interlayer cations are exchangeable, the exchange capacity of vermiculite is high, amount- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 30. 64 3 SOIL MINERALOGY Copyrighted Material Figure 3.30 Schematic diagram of the structure of chlorite. ing to 100 to 150 meq/100 g. The basal spacing, d(001), is influenced by both the type of cation and the hy-dration state. With potassium or ammonium in the exchange positions, the basal spacing is only 10.5 to 11 A˚ . Lithium gives 12.2 A˚ . Interlayer water can be driven off by heating to temperatures above 100C. This dehydration is accompanied by a reduction in basal spacing to about 10 A˚ . The mineral quickly re-hydrates and expands again to 14 A˚ when exposed to moist air at room temperature. Morphology and Surface Area Illite usually occurs as very small, flaky particles mixed with other clay and nonclay materials. High-purity deposits of illite are uncommon. The flaky par-ticles may have a hexagonal outline if well crystallized. The long axis dimension ranges from 0.1 m or less to several micrometers, and the plate thickness may be as small as 3 nm. An electron photomicrograph of illite is shown in Fig. 3.29. Vermiculite may occur in nature as large crystalline masses having a sheet structure somewhat similar in appearance to mica. In soils, ver-miculite occurs as small particles mixed with other clay minerals. The specific surface area of illite is about 65 to 100 m2 /g. The primary surface of vermiculites is 40 to 80 m2 /g, and the secondary (interlayer) surface may be as high as 870 m2/g. 3.19 OTHER CLAY MINERALS Chlorite Minerals Structure The chlorite structure consists of alter-nating micalike and brucitelike layers as shown sche-matically in Fig. 3.30. The structure is similar to that Figure 3.29 Electron photomicrograph of illite from Morris, Illinois. Picture width is 7.5 m (Tovey, 1971). of vermiculite, except that an organized octahedral sheet replaces the double water layer between mica layers. The layers are continuous in the a and b direc-tions and stacked in the c direction. The basal spacing is fixed at 14 A˚ . Isomorphous Substitution The central sheet of the mica layer is trioctahedral, with magnesium as the pre-dominant cation. There is often partial replacement of Mg2 by Al3, Fe2 and Fe3. There is substitution of Al3 for Mg2 in the brucitelike layer. The various members of the chlorite group differ in the kind and amounts of substitution and in the stacking of succes-sive layers. The cation exchange capacity of chlorites is in the range of 10 to 40 meq/100 g. Morphology Chlorite minerals occur as micro-scopic grains of platy morphology and poorly defined crystal edges in altered igneous and metamorphic rocks and their derived soils. In soils, chlorites always appear to occur in mixtures with other clay minerals. Chain Structure Clay Minerals A few clay minerals are formed from bands (double chains) of silica tetrahedra. These include attapulgite and imogolite. They have lathlike or fine threadlike morphologies, with particle diameters of 5 to 10 nm and lengths up to 4 to 5 m. An electron photomicro-graph of bundles of attapulgite particles is shown in Fig. 3.31. Although these minerals are not frequently encoun-tered, attapulgite is commercially mined and is used as a drilling mud in saline and other special environments because of its high stability in suspensions. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 31. DETERMINATION OF SOIL COMPOSITION 65 Copyrighted Material Figure 3.31 Electron photomicrograph of attapulgite from Attapulgis, Georgia. Picture width is 4.7 m (Tovey, 1971). Mixed-Layer Clays More than one type of clay mineral is usually found in most soils. Because of the great similarity in crystal structure among the different minerals, interstratifica-tion of two or more layer types often occurs within a single particle. Interstratification may be regular, with a definite repetition of the different layers in sequence, or it may be random. According to Weaver and Pollard (1973), randomly interstratified clay minerals are sec-ond only to illite in abundance. The most abundant mixed-layer material is com-posed of expanded water-bearing layers and contracted non-water-bearing layers. Montmorillonite–illite is most common, and chlorite–vermiculite and chlorite– montmorillonite are often found. Rectorite is an inter-stratified clay with high charge, micalike layers with fixed interlayer cations alternating in a regular manner with low-charge montmorillonite-like layers containing exchangeable cations capable of hydration. Noncrystalline Clay Materials Allophane Clay materials that are so poorly crys-talline that a definite structure cannot be determined are termed allophane. Such material is amorphous to X-rays because there is insufficient long-range order of the octahedral and tetrahedral units to produce sharp diffraction effects, although in some cases there may be diffraction bands. Allophane has no definite com-position or shape and may exhibit a wide range of physical properties. Some noncrystalline clay material is probably contained in all fine-grained soils. It is common in volcanic soils because of the abundance of glass particles. Oxides All soils probably contain some amount of colloidal oxides and hydrous oxides (Marshall, 1964). The oxides and hydroxides of aluminum, silicon, and iron are most frequently found. These materials may occur as gels or precipitates and coat mineral particles, or they may cement particles together. They may also occur as distinct crystalline units; for example, gibb-site, boehmite, hematite, and magnetite. Limonite and bauxite, which are noncrystalline mixtures of iron and aluminum hydroxides, are also sometimes found. Oxides are particularly common in soils formed from volcanic ash and in tropical residual soils. Some soils rich in allophane and oxides may exhibit signif-icant irreversible decreases in plasticity and increases in strength when dried. Many are susceptible to break-down and strength loss when subjected to traffic or manipulation during earthwork construction (Mitchell and Sitar, 1982; Mitchell and Coutinho, 1991). 3.20 SUMMARY OF CLAY MINERAL CHARACTERISTICS The important structural, compositional, and morpho-logical characteristics of the important clay minerals are summarized in Table 3.6. Data on the structural characteristics of the tetrahedral and octahedral sheet structures are included. 3.21 DETERMINATION OF SOIL COMPOSITION Introduction Identification of the fine-grained minerals in a soil is usually done by X-ray diffraction. Simple chemical tests can be used to indicate the presence of organic matter and other constituents. The microscope may be used to identify the constituents of the nonclay frac-tion. Accurate determination of the proportions of dif-ferent mineral, organic, and amorphous solid material in a soil, while probably possible with the expenditure of great time and at great cost, is unlikely to be worth-while owing to our inability to make exact quantitative links from composition to properties. Accordingly, from knowledge of grain size distribution, the relative intensities of different X-ray diffraction peaks, and a few other simple tests a semiquantitative analysis may be made that is usually adequate for most purposes. A general approach is given in this section for the determination of soil composition, some of the tech-niques are described briefly, and criteria for identifi-cation of important soil constituents are stated. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 32. 66 3 SOIL MINERALOGY Table 3.6 Summary of Clay Mineral Characteristics Structural 1. Silica Tetrahedron: Si atom at center. Tetrahedron units form hexagonal network Si4O8(OH)4 2. Gibbsite Sheet: Aluminum in octahedral coordination. Two-thirds of possible positions filled. Al2(OH)—O—O 2.60 A˚ . 3. Brucite Sheet: Magnesium in octahedral coordination. All possible positions filled. Mg2(OH)—O—O 2.60 A˚ . Type Subgroup and Schematic Structure Mineral Complete Formula / Unit Cella Octahedral Layer Cations Tetrahedral Layer Cations Structure Isomorphous Substitution Interlayer Bond Allophane Allophanes Amorphous — — Copyrighted Material Kaolinite Kaolinite (OH)8Si4Al4O11 Al4 Si4 Little O—OH Hydrogen Strong 11 Dickite Nacrite Halloysite (dehydrated) Halloysite (hydrated) (OH)8Si4Al4O10 (OH)8Si4Al4O10 (OH)8Si4Al4O10 (OH)8Si4Al4O10 4H2O Al4 Al4 Al4 Al4 Si4 Si4 Si4 Si4 Little Little Little Little O—OH Hydrogen Strong O—OH Hydrogen Strong O—OH Hydrogen Strong O—OH Hydrogen Strong Montmorillonite (OH)4Si8Al4O20 NH2O (Theoretical Unsubsitituted) Montmorillonite (OH)4Si8(Al3.34Mg.66O20nH2O ↓ * Na.66 Al3.34Mg.66 Si8 Mg for Al, Net charge always 0.66- / unit cell O—O Very weak expanding lattice Beidellite Nontronite (OH)4(Si7.34Al66)(Al4)O20nH2O ↓ Na.66 (OH)4(Si7.34Al.66)Fe4 3O20nH2O ↓ Na.66 Al4 Fe4 Si7.34Al.66 Si7.34Al.66 Al for Si, Net charge always 0.66- / for unit cell Fe for Al, Al for Si, Net charge always 0.66- / for unit cell O—O Very weak expanding lattice O—O Very weak expanding lattice 21 Saponite Hectorite Saponite Sauconite (OH)4Si8(Mg5.34Li.66)P20nH2O ↓ Na.66 (OH)4(Si7.34Al.66)Mg6O20nH2O ↓ Na.66 (Si6.94Al1.06)Al.66Fe.34Mg.36Zn4.80O20(OH)4 ↓ nH2O Na.66 Mg5.34Li.66 Mg, Fe3 Al.44Fe.34Mg.36Zn4.80 Si8 Si7.34Al.66 Si6.94Al1.06 Mg, Li for Al, Net charge always 0.66- / unit cell Mg for Al, Al for Si, Net charge always 0.66- / for unit cell Zn for Al O—O Very weak expanding lattice O—O Very weak expanding lattice O—O Very weak expanding lattice Hydrous Mica (Illite) Illites (K, H2O)2(Si)8(Al,Mg,Fe)4,6O20(OH)4 (Al,Mg,Fe)4-6 (Al,Si)8 Some Si always replaced by Al, Balanced by K between layers. K ions; strong Vermiculite Vermiculite (OH)4(Mg,Ca)x(Si8xAlx)(Mg.Fe)6O20.yH2O x 1 to 1.4, y 8 (Mg,Fe)6 (Si,Al)8 Al for Si not charge of 1 to 1.4 / unit cell Weak 211 Chlorite Chlorite (Several varieties known) (OH)4(SiAl)8(Mg.Fe)6O20 (21 layer) (MgAl)6(OH)12 interlayer (Mg,Fe)6(21 layer) (Mg,Al)6 interlayer (Si,Al)8 Al for Si in 21 layer Al for Mg in interlayer Chain Structure Sepiolite Attapulgite Si4O11(Mg.H2)3H2O2(H2O) (OH2)4 (OH)2Mg5Si8O20.4H2O Fe or Al for Mg Some for Al for Si Weak chains linked by 0 a Arrows indicate source of charge deficiency. Equivalent Na listed as balancing cation. Two formula units (Table 3.4) are required per unit cell. b Electron microscope data. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 33. DETERMINATION OF SOIL COMPOSITION 67 Table 3.6 (Continued) Units All bases in same plane. O—O 2.55 A˚ —Space for Si 0.55 A˚ —Thickness8 4.93 A˚ . C—C height 2.1 A˚ . OH—OH 2.94 A˚ . Space for ion 0.61 A˚ . Thickness of unit 5.05 A˚ . Dioctahedral. OH—OH 2.94 A˚ . Space for ion 0.61 A˚ . Thickness of unit 5.05 A˚ . Trioctahedral. Structure Crystal Structure Basal Spacing Shape Sizeb Cation Exchange Cap.(meq / 100 g) Specific Gravity Specific Surface m2 / g Occurrence in Soils of Engineering Interest Irregular, some-what rounded Material Copyrighted 0.05–1 Common Triclinic a 5.14, b 8.93, c 7.37 91.6, 104.8, 89.9 7.2 A˚ 6-sided flakes 0.1–4 single 0.05–2 to 3000 4000 (stacks) 3–15 2.60–2.68 10–20 Very common Monoclinic a 5.15, b 8.95, c 14.42 9648 14.4 A˚ Unit cell contains 2 unit layers 6-sided flakes 0.07–300 2.5– 1000 1–30 Rare Almost Orthorhombic a 5.15, b 8.96, c 43 9020 a 5.14 in O Plane a 5.06 in OH Plane b 8.93 in O Plane b 8.62 in OH Plane layers curve 43 A˚ 7.2 A˚ 10.1 A˚ Unit cell contains 6 unit layers Random stacking of unit cells Water layer between unit cells Rounded flakes Tubes Tubes 1 0.025– 0.15 0.07 O.D. 0.04 I.D. 1 long. 5–10 5–40 2.55–2.56 2.0–2.2 35–70 Rare Occasional Occasional 9.6A˚—Complete separation Dioctahedral Flakes (equi-dimensional) 10 A˚ up to 10 80–150 2.35–2.7 50–120 Primary 700–840 Secondary Very common 9.6A˚—Complete separation Dioctahedral Rare 9.6A˚—Complete separation Dioctahedral Laths Breadth 1 / 5 length to several unit cell 110–150 2.2–2.7 Rare 9.6A˚—Complete separation Trioctahedral To 1 unit cell breadth 0.02 0.1 17.5 Rare Trioctahedral Similar to mont. Similar to mont. 70–90 2.24–2.30 Rare Trioctahedral Brand laths 50 A˚ Thick Rare 10 A˚ Both dioctrahedral and trioctahedral Flakes 0.003–0.1 up to 10 10–40 2.6–3.0 65–100 Very common a 5.34, b 9.20 c 28.91, 9315 10.5–14 A˚ Alternating Mica and double H2O layers Similar to illite 100–150 40–80 Primary 870 Secondary Fairly common Monoclinic (Mainly) a 5.3, b 9.3 c 28.52, 978 14 A˚ Similar to illite 1 10–40 2.6–2.96 Common Monoclinic a 2 11.6, b 2 7.86 c 5.33 a0 Sin 12.9 b0 18 c0 5.2 Chain Double silica chains Flakes or fibers Laths Max, 4–5 50–100 A˚ Width 2t 20–30 20–30 2.08 Rare Occasional From Grim, R. E. (1968) Clay Mineralogy, 2d edition, McGraw-Hill, New York. Brown, G. (editor) (1961) The X-ray Identification and Crystal Structure of Clay Materials, Mineralogical Society (Clay Minerals Group), London. Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 34. 68 3 SOIL MINERALOGY Methods for Compositional Analysis Methods and techniques that may be employed for de-termination of soil composition and study of soil grains include: 1. Particle size analysis and separation 2. Various pretreatments prior to mineralogical analysis Copyrighted Material 3. Chemical analyses for free oxides, hydroxides, amorphous constituents, and organic matter 4. Petrographic microscope study of silt and sand grains 5. Electron microscope study 6. X-ray diffraction for identification of crystalline minerals 7. Thermal analysis 8. Determination of specific surface area 9. Chemical analysis for layer charge, cation exchange capacity, exchangeable cations, pH, and soluble salts 10. Staining tests for identification of clays Procedures for determination of soil composition are described in detail in publications of the American So-ciety of Agronomy. Part 1—Physical and Mineralog-ical Methods provides a set of procedures for mineralogical analyses for use by soil scientists and engineers. Part 2—Microbiological and Biochemical Properties, published in 1994, is useful for determi-nations needed for bioremediation and other geoen-vironmental purposes. Part 3—Chemical Methods, published in 1996 contains methods for characterizing soil chemical properties as well as several methods for characterizing soil chemical processes. Part 4— Physical Methods, published in 2002, is an updated version of the physical methods covered in Part 1. For each method, principles are presented as well as the details of the method. In addition, the interpretation of results is discussed, and extensive bibliographies are given. Accuracy of Compositional Analysis Techniques for chemical analysis are generally of a high order of accuracy. However, this accuracy does not extend to the overall compositional analysis of a soil in terms of components of interest in understand-ing and quantifying behavior. This is because knowl-edge of the chemical composition of a soil is of limited value by itself. Chemical analysis of the solid phase of a soil does not indicate the organization of the ele-ments into crystalline and noncrystalline components. For quantitative mineralogical analysis of the clay fraction, it is usually necessary to assume that the properties of the mineral in the soil are the same as those of a reference mineral. However, different sam-ples of any given clay mineral may exhibit significant differences in composition, surface area, particle size and shape, and cation exchange capacity. Thus, selec-tion of ‘‘standard’’ minerals for reference is arbitrary. Quantitative clay mineral determinations cannot be made to an accuracy of more than about plus or minus a few percent without exhaustive chemical and min-eralogical tests. General Scheme for Compositional Analysis A general scheme for determination of the components of a soil is given in Fig. 3.32. Techniques of the most value for qualitative and semiquantitative analysis are indicated by a double asterisk, and those of particular use for explaining unusual properties are indicated by a single asterisk. The scheme shown is by no means the only one that could be used; a feedback approach is desirable wherein the results of each test are used to plan subsequent tests. Brief discussions of the various techniques listed in Fig. 3.32 are given below. X-ray diffraction analysis is treated in more detail in the next section because of its particular usefulness for the identification of fine-grained soil minerals. Grain Size Analysis Determination of particle size and size distribution is usually done using sieve anal-ysis for the coarse fraction [sizes greater than 74 m (i.e., 200 mesh sieve)] and by sedimentation methods for the fine fraction. Details of these methods are pre-sented in standard soil mechanics texts and in the stan-dards of the American Society for Testing and Materials (ASTM). Determination of sizes by sedi-mentation is based on the application of Stokes’s law for the settling velocity of spherical particles: s w 2 v D (3.2) 18 where s unit weight of particle, w unit weight of liquid, viscosity of liquid, and D diameter of sphere. Sizes determined by Stoke’s law are not ac-tual particle diameters but, rather, equivalent spherical diameters. Gravity sedimentation is limited to particle sizes in the range of about 0.2 mm to 0.2 m, the upper bound reflecting the size limit where flow around the particles is no longer laminar, and the lower bound representing a size where Brownian motion keeps par-ticles in suspension indefinitely. The times for particles of 2, 5, and 20 m equivalent spherical diameter to fall through water a distance of 10 cm are about 8 h, 1.25 h, and 5 min, respectively, at 20C. At 30C the required times are about 6.5 h, 1 Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com
  • 35. DETERMINATION OF SOIL COMPOSITION 69 Material Copyrighted Figure 3.32 Flow sheet for compositional analysis of soils (adapted from Lambe and Martin, 1954). h, and 4 min. A centrifuge can be used for accelerating the settlement of small particles and is the most prac-tical means for extracting particles smaller than about a micrometer in size. Sedimentation methods call for treatment of a soil– water suspension with a dispersing agent and thorough mixing prior to the start of the test. This causes break-down of aggregates of soil particles, and the degree of breakdown may vary greatly with the method of prep-aration. For example, the ASTM standard method of test permits the use of either an air dispersion cup or a blender-type mixer. The amount of material less than 2 m equivalent spherical diameter may vary by as much as a factor of 2 by the two techniques. The re-lationship between the size distribution that results from laboratory preparation of the sample to that of the particles and aggregates in the natural soil is un-known. Optical and electron microscopes are sometimes used to study particle sizes and size distributions and to provide information on particle shape, aggregation, angularity, weathering, and surface texture. Pore Fluid Electrolyte The total concentration of soluble salts may be determined from the electrical conductivity of extracted pore fluid. Chemical or pho-tometric techniques may be used to determine the el-emental constituents of the extract (Rhoades, 1982). Removal of excess soluble salts by washing the sample with water or alcohol may be necessary before pro-ceeding with subsequent analysis. If they are not re- Copyright © 2005 John Wiley Sons Retrieved from: www.knovel.com