You are on page 1of 33

Introductory Concepts Final Draft, Revised 1/18/00 Page 1

IX. REFERENCES ................................................................................... 14


An Introduction to Soil Mineralogy X. LIST OF FIGURES............................................................................. 14
XI. LIST OF TABLES .............................................................................. 15
Darrell G. Schulze, Purdue University, West Lafayette

Table of Contents
I. CHEMICAL AND STRUCTURAL CLASSIFICATION OF
MINERALS ................................................................................................... 2
A. Composition of the Earth’s Crust ..................................................... 2
B. Definition of a Mineral ..................................................................... 2
C. Mineral Classification....................................................................... 2
1. Halide, Sulfate, and Carbonate Minerals ..................................... 3
2. Sulfides......................................................................................... 3
3. Oxides, Hydroxides and Oxyhydroxides...................................... 3
4. Silicates ........................................................................................ 4
II. PHYLLOSILICATE MINERALS IN SOILS ....................................... 4
A. Basic Structural Concepts................................................................. 4
1. Closest Packing of Spheres .......................................................... 4
2. Tetrahedra and Octahedra ............................................................ 4
3. Representing Tetrahedra and Octahedra ...................................... 5
4. Ionic Radii and Radius Ratios...................................................... 5
5. Tetrahedral and Octahedral Sheets............................................... 6
B. Phyllosilicate Minerals Common in Soils......................................... 7
1. The 1:1 Type Minerals ................................................................. 7
2. The 2:1 Type Minerals ................................................................. 7
C. Structural Details of Phyllosilicates.................................................. 9
III. OTHER ALUMINOSILICATE MINERALS COMMON IN SOILS 10
A. Zeolites ........................................................................................... 10
B. Allophane and Imogolite ................................................................ 10
IV. SOME CRYSTALLOGRAPHIC CONCEPTS .................................. 11
A. Periodicity in Crystals..................................................................... 11
B. The Unit Cell .................................................................................. 11
C. Miller Indices.................................................................................. 12
D. X-ray Diffraction ............................................................................ 12
1. Bragg's Law................................................................................ 12
2. The Scherrer Equation................................................................ 13
V. SUMMARY ........................................................................................ 13
VI. SUGGESTED READING .................................................................. 13
VII. QUESTIONS AND EXERCISES ...................................................... 13
VIII. ACKNOWLEDGMENTS................................................................... 14
Introductory Concepts Final Draft, Revised 1/18/00 Page 2

larger O atoms are in an approximately close-packed arrangement held


together by smaller metal atoms in the interstitial space. Most of the
An Introduction to Soil Mineralogy elements in the crust and in soils occur in minerals. Thus, the elements listed
in Table 1-1 are major constituents of the minerals discussed in this book.
Darrell G. Schulze, Purdue University, West Lafayette, Indiana
B. Definition of a Mineral
Minerals make up about one-half of the volume of most soils. They
provide physical support for plants and create the water- and air-filled pores Klein and Hurlbut (1993) define a mineral as follows: "A mineral is a
that make plant growth possible. Mineral weathering releases plant nutrients naturally occurring homogeneous solid with a definite (but not generally
that are retained by other minerals through adsorption, cation exchange, and fixed) chemical composition and a highly ordered atomic arrangement. It is
precipitation. Minerals are indicators of the amount of weathering that has usually formed by inorganic processes." Both chemical composition and
taken place and the presence or absence of particular minerals gives clues to crystal structure (ordered atomic arrangement) are important parts of this
how soils formed. The physical and chemical characteristics of soil minerals definition. Neither alone is sufficient to explain the properties of minerals.
are important considerations in planning, constructing, and maintaining Minerals with similar chemical composition but different crystal structures
buildings, roads, and airports. Soil minerals can adsorb many organic and — or, conversely, minerals with similar crystal structures but different
inorganic environmental pollutants, promoting their degradation to nontoxic chemical compositions — can be quite different from one another despite
forms, attenuating their movement through the soil, or preventing their their chemical or structural similarities.
uptake by plants and their introduction into the food chain. Some minerals
are themselves pollutants and can cause serious environmental problems C. Mineral Classification
when they are exposed to weathering at the soil surface by man’s activities.
An understanding of soil mineralogy is central to understanding virtually all Although different classification schemes could be used, mineralogists
facets of man’s use and misuse of soils and is often the key to solving have determined that first separating minerals into groups based on their
specific environmental problems. chemical composition gives classes with the greatest similarities in many
This chapter develops a core of concepts and terminology needed for other properties. Thus, minerals are first divided into classes depending upon
understanding soil minerals. The chemical composition of the earth's crust is the dominant anion or anionic group. The classes include: (1) native
discussed first to show that the most abundant elements in the crust are, not elements, (2) sulfides, (3) sulfosalts, (4) oxides and hydroxides, (5) halides,
surprisingly, the ones most likely to be encountered in soil minerals. Then (6) carbonates, (7) nitrates, (8) borates, (9) phosphates, (10) sulfates,
the chemical and structural classification of minerals is discussed and the (11) tungstates, and (12) silicates (Klein and Hurlbut, 1993). These classes
major minerals represented in soils are mentioned. The phyllosilicate are then subdivided based on chemical and structural similarities. This same
minerals are covered separately because of their major role in soils. Basic general classification is followed in this book, but with some exceptions.
structural concepts common to all minerals are covered at this point to Mineral classes such as the native elements, sulfosalts, nitrates, borates, and
provide the background for the discussion of the phyllosilicate structures. tungstates that occur only rarely in soils, are covered only briefly or not at
The overall structural theme of the phyllosilicates is then presented along all. Commonly occurring mineral classes, particularly the silicates, are
with a very brief summary of their most important properties. The chapter covered in detail.
concludes with a treatment of some crystallographic and x-ray diffraction Soil minerals are also referred to as either primary or secondary
concepts used throughout the book. minerals. Primary minerals form at elevated temperatures. They are usually
derived from igneous or metamorphic rocks, but they can be inherited from
I. CHEMICAL AND STRUCTURAL CLASSIFICATION OF sedimentary rocks as well. Secondary minerals are formed by low-
MINERALS temperature reactions and are either inherited from sedimentary rocks or
formed in soils by weathering (Jackson, 1964). The separation of minerals
into primary and secondary mineral classes is not mutually exclusive and
A. Composition of the Earth’s Crust some minerals can occur in both. The concept is useful however, and
appears widely in the soil science literature. The major mineral classes
Most of the weight and volume of the earth's crust is made up by only a represented in soils are discussed below.
few elements (Table 1-1). Oxygen and Si make up most of the weight, while
oxygen alone accounts for more than 90% of the volume. In general, the
Introductory Concepts Final Draft, Revised 1/18/00 Page 3

1. Halide, Sulfate, and Carbonate Minerals of acidity causes problems in utilizing pyrite-containing soils and in
revegetating mined areas and is a serious and costly environmental problem.
The major soil minerals of this group are halite (NaCl), gypsum
(CaSO4•2H2O), calcite (CaCO3), and dolomite [CaMg(CO3)2] (Table 1-2). 3. Oxides, Hydroxides and Oxyhydroxides
This group is characterized by minerals with relatively simple structures.
The halite structure is one of the simplest of all minerals. It consists of Primary minerals break down during weathering and release cations and
alternating Na+ and Cl- ions arranged in cubic closest packing (Fig. 1-1). The anions that recombine to form other more stable minerals. Several elements,
other minerals in this group have similar structures with cations such as in particular Al, Fe, and Mn, form oxide, hydroxide, or oxyhydroxide
Ca2+, Mg2+, or Fe2+ alternating with anions such as S22-, SO42-, or CO32-. The minerals that are stable in the soil weathering environment. Representative
bonds between the cations and anions are predominantly ionic. These mineral species are given in Table 1-2.
minerals are among the most soluble and the softest of all soil minerals and
they are easily broken down by physical and chemical weathering processes. a. Aluminum
They occur mainly in soils of arid regions or in youthful soils in more humid Gibbsite, [Al(OH)3] is the most common Al hydroxide mineral in soils.
regions, where weathering has been minimal. It is generally associated with the latter stages of weathering when leaching
Halite is the most soluble of this group and accumulates in only the most of silica has progressed to the point that phyllosilicate minerals no longer
arid environments. It is one of the minerals present in the salic horizon of form. Gibbsite is common in highly weathered Oxisols of tropical regions. It
Aridisols. Gypsum is about 100 times less soluble than halite, but it too is has a very low cation exchange capacity and contributes to the low native
abundant only in arid regions. Gypsum is a major mineral in the gypsic fertility of most Oxisols. Gibbsite is also commonly found at the weathering
horizon of Aridisols. Halite and gypsum could also occur in soils that have interface between igneous rock and saprolite and in Andisols formed from
been contaminated by salt water, or by soluble salts leaching from industrial volcanic ash. Despite the high solution silica content in these environments,
stock piles or waste piles. gibbsite precipitates as a metastable phase that eventually dissolves and
Calcite and dolomite are common carbonate minerals that occur in a gives way to more silica-rich minerals (Sposito, 1989, p. 101).
variety of soils. These minerals precipitate in the soil profile in arid and
semiarid climates (aridic and some ustic soil moisture regimes). Calcic and b. Iron
petrocalcic horizons form if the accumulation is great enough. Carbonate Iron oxide minerals form from Fe released from primary minerals. Iron
minerals in many soils are inherited from limestone or other calcareous oxides are strong pigments and small amounts of these minerals account for
parent materials. Carbonates are usually stable and can be found throughout most of the brown and red colors of soils. Goethite (FeOOH) is the most
the soil profile under aridic and ustic soil moisture regimes (arid to common mineral of this group. It accounts for the brownish to yellowish
subhumid climates), but are leached from the soil profile and are generally color of many soils, although it may be present in only small quantities.
present only in the C horizons under the udic soil moisture regime (humid Hematite (Fe2O3) is only slightly less common than goethite and usually
climates). Calcite and dolomite can be introduced into soils originally free of occurs in association with it. Hematite is usually bright red and is
these minerals via the limestone aggregate used for road construction in responsible for the red color of many soils. Goethite and hematite are stable
some areas. minerals in an oxidizing environment. Large amounts of these two minerals
in well-aerated soils, usually in association with gibbsite and kaolinite,
2. Sulfides usually indicate an advanced stage of weathering. In soils that are saturated
with water for at least some time during the year, the very insoluble Fe3+ in
Pyrite, FeS2, the most common mineral in this group, does not occur goethite, hematite, and other iron oxides can be reduced to very soluble Fe2+.
extensively in soils, but when it is present it causes some unique problems. The Fe2+ can easily move with the soil water to other parts of the soil profile,
Pyrite precipitates in soils on wet tidal flats and river deltas of some coastal or even to other associated soils in the landscape, where it can then reoxidize
areas and also occurs in some geologic formations originally deposited in to Fe3+ and reprecipitate as goethite, lepidocrocite, or ferrihydrite when
similar environments. Thus, pyrite often occurs in close association with oxidizing conditions return. Repeated cycles of oxidation and reduction give
coal. Pyrite is unstable under oxidizing conditions and weathers quickly rise to mottles and nodules that reflect an inhomogeneous distribution of
when pyritic soils are drained or when mining leaves pyritic material on the iron oxide minerals within a soil. Soil scientists use these inhomogeneous or
surface. The weathering products include the minerals jarosite, mottled color patterns to estimate the depth to a soil water table and the
KFe3(SO4)2(OH)6, and gypsum and sulfuric acid, H2SO4. The large amount length of time the soil is saturated during the year. Thus, when delineating
wetlands, determining the best location for a septic system leach field, or
Introductory Concepts Final Draft, Revised 1/18/00 Page 4

mapping soils, soil scientists rely on predictable soil color patterns that result their importance in soils and sediments, phyllosilicates are discussed in
from iron oxide mineralogy and distribution. detail below and in other chapters of this book.

c. Manganese II. PHYLLOSILICATE MINERALS IN SOILS


Manganese oxides and hydroxides (Table 1-2) are commonly found in
soils as brown or black nodules or as thin coatings on the faces of soil Phyllosilicate minerals strongly influence both the chemical and
structural units. They are often associated with Fe oxides. Manganese occurs physical properties of soils because of their generally small particle sizes,
frequently as birnessite or lithiophorite in soils. Manganese can be oxidized high surface areas, and unique cation exchange properties. A clear
and reduced in the soil environment similar to iron. Thus, the understanding of the phyllosilicate minerals is central to understanding soil
inhomogeneous distribution of Mn into nodules is an indicator of reduction clay mineralogy and many environmental processes.
and oxidation as a result of periodic water saturation. The discovery in the early 1920's that clays are crystalline was the key to
understanding many of the properties of soils and clays. The crystal
d. Titanium structures of the phyllosilicates define the different mineral species and
Rutile and ilmenite (Table 1-2) occur in soils mainly as primary account for many of their unique properties. The structural schemes of these
minerals inherited from igneous rocks. Anatase is less common and is minerals are discussed in sections II and III. The purpose is to give an
generally considered a secondary mineral. Although frequently found in overview of the most commonly occurring structures and on the major
soils, these minerals do not occur in sufficient quantity to impact soil properties of each major mineral group. Subsequent chapters will add
physical or chemical properties. details, if necessary. Emphasis will be placed on developing an
understanding of the various structural representations encountered in the
4. Silicates literature.

The silicate mineral class is an extremely large and important group of A. Basic Structural Concepts
minerals. Nearly 40% of the common minerals are silicates, as are most
minerals in igneous rocks. Silicates constitute well over 90% of the earth's The following discussion will treat atoms as rigid spheres. This is an
crust (Klein and Hurlbut, 1993) and comprise the bulk of most soils as well. oversimplified but convenient model for developing the key concepts of
Silicates occur as both primary minerals inherited from igneous or mineral structures.
metamorphic rocks and as secondary minerals formed from the weathering
products of primary minerals. 1. Closest Packing of Spheres
As explained in more detail below, the fundamental unit of all silicate
structures is the SiO4 tetrahedron. It consists of four O2- ions at the apices of The most efficient way of packing spheres in one plane is called
a regular tetrahedron coordinated to one Si4+ at the center. The individual hexagonal closest packing (Fig. 1-2). Each sphere touches six of its
tetrahedra are linked together by sharing O2- ions to form more complex neighbors and a hexagon results if the centers of these six neighbors are
structures. Several different arrangements of the SiO4 tetrahedra occur, connected. There are two types of voids between the spheres: "A" voids
partly accounting for the large number of silicate minerals and providing the where the triangular outline points downward in the figure, and "B" voids
basis for their classification. The tetrahedra may be present as single where the triangular outline points upward. If a second plane of spheres is
tetrahedra (nesosilicates), double tetrahedra (sorosilicates), rings placed on top of the first (Fig. 1-3), and if the first sphere of the second
(cyclosilicates), single or double chains (inosilicates), sheets (phyllosilicates) plane is placed in the dimple formed by the three spheres surrounding an A
or three-dimensional frameworks (tectosilicates) (Table 1-3). In most of void, then all the remaining spheres of the second plane must be placed over
these arrangements adjacent SiO4 tetrahedra share corners, that is, they share A voids as well. The spheres illustrated in Fig. 1-3 represent O2-, OH-, or F-
a common O2-. The more common minerals from each silicate class likely to in the phyllosilicate structures, while positively charged cations occupy the
be found in soils are given in Table 1-3. space between these negatively charged anions.
The most common minerals in igneous rocks are the olivines,
pyroxenes, amphiboles, micas, feldspars, and quartz. These primary minerals 2. Tetrahedra and Octahedra
predominate in the sand and silt size fractions of soils. The clay (<2 µm)
fraction of most soils is dominated by phyllosilicate minerals. Because of Another examination of Fig. 1-3 shows that there are two different types
of sites into which cations can be placed. The "A" sites, formed over the "A"
Introductory Concepts Final Draft, Revised 1/18/00 Page 5

voids of the lower plane of spheres, are called tetrahedral sites because the the oxygens and hydroxyls at the corners of the polyhedra (see, for example,
four O2- ions surrounding this site form the apices of a regular tetrahedron. Fig. 1-9 later in this chapter), while other models show only sticks to
The "B" sites, formed over the "B" voids of the lower plane of spheres, are represent bonds, but no spheres to represent the atoms.
called octahedral sites because the six O2- ions surrounding each of these There is some ambiguity associated with each type of representation,
sites form the apices of a regular octrahedron (Fig. 1-3). This illustrates that and it is important to understand the correspondence between, and
both octahedral and tetrahedral sites are a consequence of the closest limitations of, each of the different structural representations. Thus, it is
packing of two planes of spheres. important to utilize as many different learning aids as possible. Computer
A cation in a tetrahedral site is in 4-fold (or tetrahedral) coordination graphics programs that allow the viewer to rotate and manipulate structural
because it is bonded to four O2- ions whose centers define the apices models on the computer screen offer significant advantages over static
(corners) of a tetrahedron. Likewise, a cation in an octahedral site is in drawings. Physical models should be studied as well, if they are available.
6-fold (or octahedral) coordination because it is bonded to six O2- ions Far from being child’s play, building your own models is one of the best
whose centers define the apices of an octahedron. Tetrahedrally and ways to learn about mineral structures. Models can be built with inexpensive
octahedrally coordinated cations—tetrahedra and octahedra for short—are balls and some glue or with paper polyhdera [see Moore and Reynolds
common structural elements in many mineral structures. The approach of (1997) for templates].
describing mineral structures as assemblages of polyhedra is used widely
and, once mastered, provides an efficient way to understand even complex 4. Ionic Radii and Radius Ratios
mineral structures.
Whether or not a given atom can occupy a tetrahedral or an octahedral
3. Representing Tetrahedra and Octahedra site depends both on the atom’s charge and size. The size of atoms and ions
varies depending on their atomic number, ionization state, and coordination
Octahedra and tetrahedra are commonly represented using three number (Table 1-4).
different types of models. Each model portrays the same concept, but The central void within a tetrahedron is smaller than the one within an
different representations are necessary to show various structural features. octahedron. In other words, the largest sphere that can be placed in a
Fig. 1-3 shows two views of a tetrahedron and of an octahedron drawn in tetrahedral site without pushing the O2- ions apart is smaller than the sphere
these three different ways. that can be placed in an octahedral site without pushing. Simple calculations
The sphere-packing model gives an impression of the space occupied by (see, for example, Klein and Hurlbut, 1993) illustrate that the maximum
the atoms in a structure. The sphere-packing model is particularly useful for radius of a sphere that will just fit in the tetrahedral site is 0.225 times the
visualizing the shape of the external surface of a mineral, but it does not radii of the 4 surrounding O2- ions. For an octahedral site, the maximum
allow one to easily see the interior of the structure. The ball-and-stick model radius is 0.414 times the radii of the 6 surrounding O2- ions. The next larger
shows the bonds within the crystal since the bonds are represented as type of site (not illustrated) is a cubic site consisting of 8 O2- ions arranged at
"sticks" connecting the balls. The tetrahedra and octahedra are more difficult the corners of a cube. For a cubic site, the maximum radius is 0.732 times
to visualize, but the interior of the structure is easier to see because the the radii of the 8 surrounding O2- ions. Assuming a radius of 0.140 nm for
atoms are drawn much smaller than in the close-packed model. The the O2- ions, the limiting radius for a tetrahedral site is 0.032 nm (0.140 nm
polyhedral models give the best impression of tetrahedra and octahedra since × 0.225 = 0.032 nm), for an octahedral site it is 0.058 nm, and for a cubic
the atoms are represented only as points in space or as small spheres. The site it is 0.102 nm. The Si4+ ion has a radius <0.032 nm (Table 1-4). Thus,
centers of the O2- ions are connected by lines to form the edges of tetrahedra Si4+ always occurs in tetrahedral rather than octahedral sites. The Mg2+ ion
and octahedra. Thus, each apex represents the position of an O2- ion, while a has a radius that is >>0.032 so Mg2+ is unlikely to occur in tetrahedral
cation resides in the center of the tetrahedron or octahedron. In general, only coordination, but since its radius is between 0.058 and 0.102 nm, it fits
those tetrahedra or octahedra actually occupied by cations are represented by easily in an octahedral site. The Al3+ ion, with a radius near to the limiting
a solid polyhedron, sites that are not occupied by cations are not represented radius of 0.032 nm, can fit in both tetrahedral and octahedral sites. Cations
by solid polyhedra. One should use care in interpreting polyhedral models, with radii >~ 0.102 nm tend to occur in cubic, dodecahedral (12-fold) or
since it is easy to get the impression that a structure contains a large amount higher coordination sites rather than in octahedral or tetrahedral sites
of open space or pores. The same structure drawn as a close-packed model (Table 1-4).
will show that much of the space is actually occupied by the large oxygen
atoms. Other ways of representing mineral structures are encountered as
well. Sometimes, polyhedral models are drawn without spheres to represent
Introductory Concepts Final Draft, Revised 1/18/00 Page 6

5. Tetrahedral and Octahedral Sheets The discussion above followed a sphere-packing approach in which we
showed that the structure of the octahedral sheets could be derived assuming
Tetrahedra arranged into sheets are common to the structures of all that the large oxygen atoms were hard spheres of identical size closely
phyllosilicates. Octahedra arranged into sheets are present in the structures packed in three dimensions. Thus, we assumed that the oxygen atoms were
of phyllosilicates and in some hydroxide minerals as well. The structure of our basic structural units and that the cations merely served to balance the
tetrahedral and octahedral sheets will be discussed first, then it will be negative charge of the oxygens. This approach works for some of the oxide
shown how different combinations of tetrahedral and octahedral sheets give and hydroxide minerals, but in many mineral structures the oxygen atoms
rise to the different clay mineral structures. Most of the diagrams in this are not as closely packed as strict geometric packing would dictate. A more
chapter show ideal crystal structures based on the approximate closest general way of describing mineral structures is called the polyhedral
packing of spheres. approach. In the polyhedral approach, the coordination polyhedra, primarily
octahedra and tetrahedra in the phyllosilicates, are considered the basic
a. The Octahedral Sheet structural building blocks. Thus, we can consider the octahedral sheets
Consider two planes of spheres representing hydroxyl (OH-) ions in illustrated in Fig. 1-4 as an assemblage of octahedra in which adjacent
hexagonal closest packing. (The H+ takes up very little space and the OH- octahedra share two oxygens with one another. In other words, the octahedra
can be considered a sphere of roughly the same size as an O2- ion). There are share edges with one another. In general, coordination polyhedra can share
two ways of filling the octahedral sites depending on the valence of the corners (one shared oxygen), edges (two shared oxygens), or faces (three or
cation. A divalent cation, such as Mg2+, can be placed into each octahedral more shared oxygens) with adjacent polyhedra. The octahedral sheet could
site to obtain the network of octahedra illustrated in Fig. 1-4a. This be described using either the sphere packing or polyhedral approaches. The
arrangement is called trioctahedral because 3 of every 3 octahedral sites are tetrahedral sheet described below can only be describe using the polyhedral
occupied by a cation. This gives a formula of Mg3(OH)6 or Mg(OH)2 for the approach because the oxygens in the tetrahedral sheet are not close-packed
whole sheet; each Mg2+ is surrounded by 6 OH-'s, but since each OH- is as they are in the octahedral sheet.
shared equally among 3 different Mg2+ atoms, each OH- contributes only 1/3
of its negative charge to each Mg2+, giving the formula Mg(OH)2. The sheet b. The Tetrahedral Sheet
is electrically neutral because the charge is balanced within the sheet. The tetrahedral sheet consists of SiO4 tetrahedra arranged such that
The other possibility is to place a trivalent cation, such as Al3+, into the three O2- ions of each tetrahedron are shared with the three nearest neighbor
octahedral sites. To preserve charge neutrality, only 2 trivalent cations are tetrahedra (Fig. 1-5). These shared O2- ions are all in the same plane and are
needed compared to 3 divalent cations, giving the formula Al2(OH)6 or referred to as basal oxygens. Note that two adjacent tetrahedra share only
Al(OH)3. Thus, only 2 of every 3 possible octahedral sites are filled. This one O2- between them. (The tetrahedra share apices or corners.) The fourth
arrangement is called dioctahedral and is illustrated in Fig. 1-4b. The unit O2- ion of each tetrahedron is not shared with another SiO4 tetrahedron and
formula for the dioctahedral sheet can be deduced by following the same is free to bond to other polyhedral elements. These unshared O2- ions are
reasoning used above. Each Al3+ is surrounded by six OH-'s, but since each referred to as apical oxygens. Since each basal oxygen contributes a formal
OH- is shared equally between 2 different Al3+ ions (not 3 as for Mg2+), each charge of -1 to each Si4+ ion, the addition of H+ ions to the apical oxygens to
OH- contributes 1/2 of its negative charge to each Al3+, giving the formula form hydroxyls should result in an electrically neutral tetrahedral sheet. Such
Al(OH)3. Again the charge is balanced and the sheet is electrically neutral. individual tetrahedral sheets, although sometimes postulated to occur as
Note how the trioctahedral and dioctahedral sheets appear in the transient weathering products in aqueous solution, do not stack to form
different types of representations. The sphere-packing model shows that stable mineral structures on their own. Thus, tetrahedral sheets only occur in
both the trioctahedral and dioctahedral sheets contain the same dense combination with octahedral sheets as described below.
packing of OH-'s. The pattern of unoccupied octahedral sites is more evident The upper half of Fig. 1-5 shows all of the apical oxygens pointing in
in the polyhedral model, where unoccupied octahedra appear as open spaces, the same direction, namely, out of the plane of the paper towards the reader.
and in the ball-and-stick model. Structures in which the apical oxygens of a single sheet all point in the same
Octahedral sheets, stacked one on top of the other and held together by direction are the most common, but structures also occur in which the apical
weak residual bonds, make up the structures of gibbsite (Fig. 1-7) and oxygens point alternately in opposite directions. Minerals containing this
brucite [Mg(OH)2]. Gibbsite and brucite differ in that gibbsite contains Al in
the octahedral sites and is dioctahedral, while brucite contains Mg in the
octahedral sites and is trioctahedral. The structures of gibbsite and brucite
are the simplest in a series of structures containing octahedral sheets.
Introductory Concepts Final Draft, Revised 1/18/00 Page 7

sheet-like arrangement of SiO4 tetrahedra are called "phyllosilicates1" or Kaolinite is a common mineral in soils and is the most common member
"sheet silicates.” of this subgroup. It tends to be particularly abundant in more weathered soils
such as Ultisols and Oxisols. Kaolinites have very little isomorphous
B. Phyllosilicate Minerals Common in Soils substitution in either the tetrahedral or octahedral sheets and most kaolinites
are close to the ideal formula Al2Si2O5(OH)4. The 1:1 layer has little or no
Phyllosilicates are divided into two groups, 1:1 and 2:1-type minerals, permanent charge because of the low amount of substitution. Consequently,
based on the number of tetrahedral and octahedral sheets in the layer cation exchange capacities and surface areas are typically low. Soils high in
structure. kaolinite are generally less fertile than soils in which 2:1 clay minerals
dominate.
1. The 1:1 Type Minerals Kaolinite can form in soils from Al and Si released by the weathering of
primary and other secondary minerals. For example, feldspars often weather
to kaolinite in soils formed from igneous rocks. Kaolinite can also be
a. The 1:1 Layer Structure. inherited from the soil parent material — i.e. clayey sediments. Kaolinite is
The 1:1 layer structure consists of a unit made up of one octahedral and used widely in a variety of industrial applications such as fillers for plastics,
one tetrahedral sheet, with the apical O2- ions of the tetrahedral sheets being in ceramics, and as a filler and coating for paper.
shared with (and part of) the octahedral sheet (Fig. 1-6a and 1-6b). There are
three planes of anions (Fig. 1-6b). One plane consists of the basal O2- ions of c. Halloysite.
the tetrahedral sheet, the second consists of O2- ions common to both the Halloysite has a 1:1 layer structure similar to kaolinite except that the
tetrahedral and octahedral sheets plus OH- belonging to the octahedral sheet, 1:1 layers are separated by a layer of H2O molecules when fully hydrated
and the third consists only of OH- belonging to the octahedral sheet.2 (Fig. 1-7). Most clay silicates occur as thin plates, but halloysite often occurs
The edge view of the 1:1 layer in Fig. 1-6b is one of two possible as tubular or spherical particles.
depictions often encountered in the literature. The other view is obtained by Halloysite is usually found in soils formed from volcanic deposits,
rotating the structure around an axis normal to the sheet and viewing the particularly volcanic ash and glass. It is a common clay mineral in the
structure along a different crystallographic axis (Fig. 1-6c). Andisol soil order. Halloysite forms early in the weathering process but it is
generally less stable than kaolinite and gives way to kaolinite with time.
b. Kaolinite.
The structure of kaolinite consists of 1:1 layers stacked one above the 2. The 2:1 Type Minerals
other. Kaolinite is dioctahedral and contains Al3+ in the octahedral sites and
Si4+ in the tetrahedral sites (Fig. 1-7). The 1:1 layer is electrically neutral and In contrast to the 1:1 minerals, which are mostly represented in soils by
adjacent layers are held together by hydrogen bonding between the basal only 2 major minerals, the 2:1 minerals are structurally more diverse and are
oxygens of the tetrahedral sheet and the hydroxyls of the surface plane of the represented by several mineral species.
adjacent octahedral sheet.
a. The 2:1 Layer Structure.
The 2:1 layer structure consists of 2 tetrahedral sheets with one bound to
1
The prefix phyllo-, derived from the Greek word phyllon, meaning leaf, should not be
each side of an octahedral sheet (Fig. 1-6b). There are four planes of anions.
confused with the prefix phylo-, derived from the Greek word phylon, meaning tribe. The outer two consist of the basal oxygens (O’s) of the two tetrahedral
Phyllo- is pronounced fil’o, while phylo- is pronounced fi’ lo. Thus, the proper sheets, while the two inner planes consist of oxygens common to the
pronunciation for phyllosilicate is fil’o sil’ i kat. octahedral sheet and one of the tetrahedral sheets, plus the hydroxyls (OH’s)
2 belonging to the octahedral sheet.
The usage of the terms plane, sheet, and layer follows the recommendations of the
AIPEA Nomenclature Committee. They refer to increasingly thicker arrangements such
as a single plane of atoms, a tetrahedral or octahedral sheet and a 1:1 or 2:1 layer. A b. Talc and Pyrophyllite.
sheet is a combination of planes, and a layer is a combination of sheets. Layers can be The simple structures of talc and pyrophyllite are good starting points
separated by various interlayer materials, such as cations, hydrated cations, organic for discussing 2:1 structures. Both minerals consist of 2:1 layers stacked one
molecules, and hydroxide octahedral groups and sheets. The total assembly of a layer
plus interlayer material is called a unit structure (AIPEA Nomenclature Committee, above the other. Talc has Mg2+ in the octahedral sites and is trioctahedral,
1980). while pyrophyllite (Fig. 1-7) has Al3+ in the octahedral sites and is
dioctahedral. The tetrahedral sheets of both minerals contain only Si4+,
Introductory Concepts Final Draft, Revised 1/18/00 Page 8

giving ideal formulas of Mg3Si4O10(OH)2 for talc and Al2Si4O10(OH)2 for exchangeable cations, primarily Ca and Mg, in the interlayer (Fig. 1-7). The
pyrophyllite. In each case the charge is balanced within the 2:1 layer, high charge per formula unit gives vermiculites a high CEC and causes them
making it electrically neutral. Adjacent 2:1 layers are held together only by to have a high affinity for weakly hydrated cations such as K+, NH4+ and
weak van der Waals forces. Cs+. Fixation of K+ by vermiculites can be significant in soils high in
Talc and pyrophyllite occur only rarely in soils, usually only when they vermiculite.
are inherited from low-grade metamorphic rocks. Talc and pyrophyllite are Vermiculites in soils are believed to form almost exclusively from the
used industrially as ingredients in paints, ceramics, plastics, paper, and weathering of micas and chlorites. The weathering of micas to vermiculite
cosmetics. The occurrence of talc in river and estuarial sediments when there (or smectite) is believed to occur by replacement of K+ in the interlayer sites
is no geological source is a reflection of industrial activity in the watershed. with hydrated exchangeable cations. The integrity of the 2:1 layer is
preserved but there is a reduction in the layer charge. Vermiculite does not
c. Micas. swell as extensively as smectite and this is illustrated in Fig. 1-7 by the
Mica minerals have the 2:1 layer structure described for talc and presence of only two planes of water molecules surrounding the hydrated
pyrophyllite but with two important differences. First, instead of having only cations in the interlayer space. Some commercial uses of vermiculites
Si4+ in the tetrahedral sites, one-fourth of the tetrahedral sites are occupied include horticultural potting media and thermal insulating materials.
by Al3+. Because of this substitution, there is an excess of one negative
charge per formula unit in the 2:1 layer. Second, this excess negative charge e. Smectites.
is balanced by monovalent cations, commonly K+, that occupy interlayer The smectite group consists of minerals with the 2:1 structure already
sites between two 2:1 layers. This gives an ideal formula of discussed for mica and vermiculite, but with a still lower charge per formula
KAl2(AlSi3)O10(OH)2 for a mica mineral with Al in the octahedral sites. weight, namely 0.6 to 0.25. As in vermiculite, the interlayer contains
Just as in talc and pyrophyllite, the octahedral sheet can be either exchangeable cations (Fig. 1-7).
dioctahedral (Fig. 1-7) or trioctahedral. There are several different mica As for the micas and vermiculites, dioctahedral smectites are more
species because Fe2+ and Fe3+ can substitute for Mg2+ and Al3+ in the common in soils than trioctahedral smectites. The most common smectite
octahedral sheet and Na+ and Ca2+ can substitute for K+ in the interlayer. minerals range in composition between the three end-members:
Mica in soils is usually inherited from the parent rock and is likely to montmorillonite, beidellite, and nontronite. All are dioctahedral, but they
occur in soils derived from various igneous and metamorphic rocks, as well differ in the composition of the tetrahedral and octahedral sheets. Smectites
as from sediments derived from them. Muscovite, biotite, and phlogopite are do not fix K+ as readily as do vermiculites because smectites have a lower
the three most common mica group minerals in rocks, and consequently in layer charge, but smectites swell more extensively than vermiculite. This is
soils. All three contain K in the interlayer (Table 1-3), but they differ in the illustrated in Fig. 1-7 by the larger spacing between the 2:1 layers.
composition of the octahedral sheet and whether they are di- or trioctahedral. Smectites are important minerals in temperate region soils because of
Mica in the clay fraction of soils and sediments differs somewhat from their high surface area and their adsorptive properties. Smectites shrink upon
the macroscopic muscovite mica it most closely resembles. This clay-size drying and swell upon wetting. This shrink-swell behavior is most
mica is often referred to as illite. Glauconite is another mica mineral that is pronounced in the Vertisol order and in vertic subgroups of other soil orders.
similar to illite, but it contains more Fe and less Al in its octahedral sheet The shrink-swell properties lead to cracking and shifting problems when
than illite. houses, roads, and other structures are built on smectitic soils. Smectites are
Micas weather to other minerals, particularly to vermiculites and used widely in industry as catalysts, adsorbents for spills, as sealants for
smectites and the K+ released during weathering is an important source of K ponds and abandoned wells, in drilling fluids for oil wells, and in liners for
for plants. As a rule, the dioctahedral micas, such as muscovite, are more landfills.
resistant to weathering than trioctahedral micas. Thus, muscovite tends to be
the most common mica mineral found in soils. Micas are used commercially f. Chlorites.
in paints and cosmetics. Like mica, chlorite minerals have a 2:1 layer structure with an excess of
negative charge. In contrast to mica, however, the excess charge is balanced
d. Vermiculites. by a positively charged interlayer hydroxide sheet (Fig. 1-7), rather than K+.
Vermiculite has a 2:1 layer structure as described for mica, but instead The interlayer hydroxide sheet is an octahedral sheet as illustrated in Fig. 1-4
of having a layer charge of ~1 per formula unit and K+ in interlayer and can be either di- or trioctahedral. Instead of being electrically neutral as
positions, vermiculite has a layer charge of 0.9 to 0.6 and contains in brucite or gibbsite, the hydroxide sheet has a positive charge cause by
Introductory Concepts Final Draft, Revised 1/18/00 Page 9

substitution of higher valence cations for lower valence ones, for example, chlorite-vermiculite, chlorite-smectite, chlorite-swelling chlorite, and
Mg2Al(OH)6+. Either octahedral sheet — the one that is part of the 2:1 layer kaolinite-smectite. Three-component mixed layer systems can also occur.
or the interlayer hydroxide sheet — can be di- or trioctahedral, and can The sequence of layers can be either regular or random. A regularly
contain Mg2+, Fe2+, Mn2+, Ni2+, Al3+, Fe3+, and Cr3+, giving a large number interstratified mineral consisting of two types of layers denoted by A and B
of different mineral species. could have a sequence like ABABAB…, or ABBABBABB…, or any other
Chlorite minerals in soils may be primary minerals inherited from either repeating sequence. In a randomly interstratified mineral the sequence of
metamorphic or igneous rocks. They may also be inherited from sedimentary layers is random — for example, ABBABAABBAAA…. Random
rocks such as shales, or from hydrothermally altered sediments. Chlorites are interstratification of layer-silicates is more common in soils than regular
rather infrequent minerals in soils and when they do occur they generally interstratification, though regular interstratification, especially in weathering
make up only a small amount of the soil. Chlorite weathers to form micas, is not rare.
vermiculite and smectite and the ease with which they break down makes Partial removal of interlayer K from micas or of interlayer hydroxide
them sensitive indicators of weathering. from chlorite is one way that interstratified minerals can form in soils. Other
possibilities include (i) fixation of adsorbed K+ by some vermiculite layers
g. Hydroxy-interlayered vermiculite and smectite. to give mica-like layers, and (ii) the formation of hydroxide interlayers to
Hydroxy-interlayered vermiculite and smectite can be considered a solid produce chlorite-like layers.
solution with vermiculite or smectite as one end member and chlorite as the
other. Hydroxy-interlayered minerals form as Al3+ released during i. Palygorskite and sepiolite.
weathering hydrolyzes and polymerizes to form large polycations with a Palygorskite and sepiolite are considered phyllosilicates, but are distinct
postulated formula of Al6(OH)153+ (or similar) in the interlayers of structurally from the typical 1:1 and 2:1 layer structures. Both minerals have
vermiculite and smectite. These polycations balance some of the charge of continuous tetrahedral sheets, but adjacent bands of tetrahedra within one
the 2:1 layer. The combination of a 2:1 layer with hydroxy Al in the tetrahedral sheet point in two opposite directions rather than in one direction
interlayer gives a structure similar to that of chlorite (Fig. 1-7). Thus, these as in the 1:1 and 2:1 structures. The result is a structure that can be described
minerals are also called secondary chlorites. The degree of filling of the as ribbons of 2:1 layers joined at their edges as illustrated in Fig. 1-8. Water
interlayer with hydroxy Al can vary from none to almost complete, with molecules occur in the spaces between the ribbons. The 2:1 ribbons are
properties of the clay varying accordingly. The interlayer hydroxy Al is not wider in sepiolite than in palygorskite.
exchangeable, therefore it lowers the CEC of smectite or vermiculite almost Palygorskite and sepiolite are often found in soils of arid and semiarid
linearly as a function of the amount of Al adsorbed in the interlayer. environments. Both minerals have a fibrous morphology in contrast to the
Interlayer hydroxy Al prevents smectite from shrinking and swelling as platy morphology of most 1:1 and 2:1 minerals. Because of their fibrous
it normally would. In vermiculite, it reduces K+ fixation by lowering the morphology, suspensions of these clays can form thick gels even at low solid
exchange capacity and by preventing the interlayer from collapsing around concentrations. Thus, palygorskite and sepiolite are used industrially as
the K+. The positively charged hydroxy interlayers also provide potential gelling agents to keep other solids in suspension.
sites for anion adsorption.
Hydroxy-interlayered vermiculite and smectite are most common in C. Structural Details of Phyllosilicates
Alfisols and Ultisols. Within a given profile they tend to be most abundant
near the soil surface. Hydroxy-interlayered smectites produced industrially Figures 1-1 through 1-8 illustrate “ideal” mineral structures based on the
are used as catalysts in the chemical industry. regular close packing of spheres. As a consequence, all of the octahedra in
Fig. 1-4 are exactly the same size and all are regular (undistorted) solids.
h. Interstratification in Phyllosilicates. Likewise, in Fig. 1-5, all of the tetrahedra are regular solids, the 6-member
Phyllosilicates in soils do not always occur as discrete particles of mica, tetrahedral rings form a regular hexagon, and the basal oxygens all lie in
vermiculite, smectite, chlorite, or kaolinite. For example, instead of being exactly the same plane. Crystallographic structure refinements of real
made up of a stack of identical 2:1 vermiculite layers, one physically discrete phyllosilicate minerals show distortions from the ideal structures illustrated
particle may consist of a mixture of both mica and vermiculite layers instead. up until now in this chapter. These distortions are similar for all of the
Such minerals are referred to as mixed-layer or interstratified minerals. phyllosilicate minerals, although they vary somewhat in magnitude.
Different types of interstratified minerals have been identified. Two- There are two reasons why real minerals deviate from these ideal
component systems include: mica-vermiculite, mica-smectite, mica-chlorite, structures. First, atoms are not hard spheres whose closest approach is
determined solely by the sum of their radii. Second, if one does try to join
Introductory Concepts Final Draft, Revised 1/18/00 Page 10

octahedral and tetrahedral sheets built of hard spheres like those illustrated phyllosilicates are more common in soils than trioctahedral phyllosilicates,
in Fig. 1-4 and 1-5, geometry alone dictates that they will not fit together if water, ions and molecules interact predominately with surfaces like those
the spheres representing the oxygen atoms touch each other in both the illustrated in Fig. 1-9, rather than surfaces with the more regular
octahedral and tetrahedral sheets. The reason becomes apparent by arrangement illustrated in Fig. 1-4 and 1-6a.
examining the sphere-packing model of the tetrahedral sheet in Fig. 5. Since
the apical oxygens are common to both the tetrahedral and octahedral sheets, III. OTHER ALUMINOSILICATE MINERALS COMMON IN
two adjacent apical oxygens in the tetrahedral sheet also define the edge of SOILS
an octahedron in the octahedral sheet (Fig. 1-6a). Note that, although basal
oxygens touch neighboring basal oxygens, apical oxygens do not touch
neighboring apical oxygens (Fig. 5). Thus, the octahedral sheets illustrated A. Zeolites
in Fig. 1-4 cannot share apical oxygens with the tetrahedral sheet illustrated
in Fig. 1-5 because the tetrahedral sheet is larger than the octahedral sheet. Zeolites are of a large group of minerals that consist structurally of SiO4
This problem was overcome in Fig. 1-6 to 1-8 by increasing the oxygen- tetrahedra arranged in ways that result in large amounts of pore space within
oxygen distance in the octahedral sheet to match the oxygen-oxygen the crystals (Fig. 1-8). Aluminum substitutes for Si in the tetrahedral sites
distance of the apical oxygens of the tetrahedral sheet. Thus, the octahedra and, as a result, the (Si,Al)O4 framework has a net negative charge. The
remain undistorted but they are larger than they would be if the spheres charge is balanced by cations that reside in the channels and pores, along
representing oxygen touched. with water molecules. Because the cations are exchangeable, zeolites have
In the real phyllosilicate structures, this mismatch between the cation exchange properties similar to the phyllosilicates, but because the
octahedral and tetrahedral sheets is compensated, not by an expansion of the tetrahedral framework of the zeolites is rigid and the size of the pores is
octahedral sheet, but primarily by distortions in the tetrahedral sheet. The net fixed, small cations can move into and out of the pores freely, while larger
effect is that, the tetrahedra are rotated from the ideal arrangement shown in cations are excluded. Thus, zeolites are often referred to as “molecular
Fig. 1-5 such that the hexagonal shapes outlined by the 6-membered sieves” because of their very selective cation exchange properties.
tetrahedral rings in Fig. 1-5 become the ditrigonal shapes illustrated in Fig. Zeolites are relatively rare in soils because they weather easily in humid
1-9. In addition, the tetrahedra are tilted such that the basal oxygens no regions, but they do occur in some soils in arid regions. Zeolites are widely
longer lie in the same plane, resulting in a corrugated basal plane in which used in industry and agriculture, for example, as catalysts and adsorbents in
some rows of atoms lie slightly below and some lie slightly above the the chemical industry, as additives in animal feeds, and as ingredients in
average plane of the basal oxygens. These corrugations are best seen in the laundry detergents. Thus, zeolites may be introduced into soils and
lowermost edge view in Fig. 1-9 where spheres representing the oxygen sediments in unexpected ways.
atoms have not been drawn. The implications of these distortions are that
molecules interacting with phyllosilicate surfaces “see”, not the ideal surface B. Allophane and Imogolite
illustrated in Fig. 1-6a, but the more distorted surface illustrated in Fig. 1-9.
The octahedral sheet is distorted from the ideal structure as well. First, The aluminosilicate minerals discussed above have three-dimensional
because of the attraction of each central octahedral cation to its 6 crystal structures with atoms packed together in a more or less regular
surrounding oxygens, octahedra occupied by cations are smaller than manner over relatively long distances (10's of nm's). They exhibit long range
octahedra in which the central sites are unoccupied (Fig. 1-9). In addition, order. Two other aluminosilicates, allophane and imogolite, exhibit short-
when two octahedra share two adjacent oxygens (the octahedra share edges), range (or local) order. Structures with short-range order exhibit order over
the oxygen-oxygen distance along this shared edge is shorter than the several nm’s, but on a larger scale the structure is disordered.
oxygen-oxygen distance along an edge that is not shared with an adjacent Allophane is a material consisting chemically of variable amounts of
occupied octahedron (Fig. 1-9). Thus, the octahedra are distorted from the O2-, OH-, Al3+, and Si4+, and characterized by short-range order and a
ideal Euclidean solids in very predictable ways. In general, trioctahedral predominance of Si-O-Al bonds. It consists of small (3.5-5.0 nm) spheres,
phyllosilicates show less structural distortion and more closely resemble the the structure of which has not been determined. The spheres clump together
“ideal” structures than do the dioctahedral phyllosilicates. to form irregular aggregates.
The structural distortions just discussed probably account for relatively Imogolite consists of tubes several µm long with an outer diameter of
minor differences in the properties of the phyllosilicates, with the major 2.3 to 2.7 nm and an inner diameter of ~l.0 nm. The tubes consist of a single
differences determined mainly by whether the mineral has a 1:1 or 2:1 layer dioctahedral sheet with the inner surface OH replaced by SiO3OH groups
structure and by the layer charge. On the other hand, since dioctahedral (Fig. 1-8; Farmer, et al., 1983). Several individual tubes are arranged in
Introductory Concepts Final Draft, Revised 1/18/00 Page 11

bundles 10-30 nm across to give thread-like particles several micrometers with three parameters. Along the X axis the pattern is repeated at intervals of
long. a. We also say that the pattern is translated (moved) along X at increments of
Allophane and imogolite usually occur as weathering products of a. The translation interval along the Y axis is designated b and the angle
volcanic ash and are important minerals in the Andisol soil order. Imogolite between the X and Y axes is designated by the angle γ. The array of leaves in
has also been identified in the Bs horizons of Spodosols. Allophane and Fig. 1-10a consists of real objects, namely, leaves. We can represent the
imogolite can specifically adsorb many inorganic and organic compounds. periodic nature of this pattern by an array of points in space, each point
Andisols, for example, usually fix large amounts of phosphate, making it having identical surroundings (Fig. 1-10b.). Such an array is called a lattice.
unavailable to plants, and the large amounts of organic matter common in A lattice of points is imaginary since each point is an imaginary infinitesimal
Andisols may be due, in part, to adsorption of organic molecules by spot in space. The concept of a lattice is useful because it allows us to
allophane and imogolite. Soils containing large amounts of allophane and represent the periodic nature of a pattern regardless of the actual object or
imogolite usually have unique physical properties such as a low bulk configuration of objects at each lattice point. The lattice depicted in Fig. 1-
density, high water holding capacity, high liquid and plasticity limits, and a 10b is called a plane lattice because it has only two dimensions. The concept
thixotropic consistence. can be extended to three dimensions to give a space lattice. A space lattice is
a regular and unlimited distribution of points in space (Fig. 1-11). The
IV. SOME CRYSTALLOGRAPHIC CONCEPTS concept of a space lattice is used in describing crystal structures because
crystals are three dimensional objects3. The directions of the crystallographic
In 1912 Max von Laue conducted an experiment to see if the "x-rays" axes in three dimensions are designated by X, Y and Z, and a, b, and c are
discovered 17 years earlier by Wilhelm Conrad Röntgen could be diffracted used for the repeat distances along these axes (AIPEA Nomenclature
by a crystal. This one experiment had two very important results: (1) it Committee, 1980). The angles between the axes are designated α between Y
proved that x-rays behaved as waves, and (2) it showed that crystals were and Z, β between X and Z and γ between X and Y.
made up of a regular array of atoms in space, and could therefore serve as a
grating to diffract the x-rays. Within a few years of von Laue's discovery, B. The Unit Cell
much of the mathematical theory describing the diffraction of x-rays by
crystals had been developed and the atomic structures of some simple Joining the points of a space lattice produces a series of parallel-sided
crystals had been determined (Azároff, 1968; Klug and Alexander, 1974). unit cells (Fig. 1-11). Each unit cell contains a complete unit of the crystal
As early as 1923 and 1924 several clays were shown to be crystalline by x- pattern because the complete pattern is reproduced at each lattice point. Note
ray diffraction, but it was not until 1930 that some of the first structure that there are alternative ways of outlining the unit cell. The choice of the
determinations of clay minerals were made. About the same time, two "best" unit cell for a mineral is made when the structure is determined. A
independent papers, one by S. B. Hendricks and W. H. Fry in 1930 and the unit cell is chosen for convenience in visualizing the symmetry and in
other by W. P. Kelly, W. H. Dore, and S. M. Brown in 1931, gave the first making mathematical calculations. Different researchers will occasionally
x-ray diffraction evidence that soil materials, even those in the finest choose different unit cells to describe the same mineral structure.
fractions, were crystalline (Marshall, 1949; Grim, 1968). There are 14 unique ways of arranging points in three-dimensional
X-ray diffraction remains the most important method for identifying soil space. These are known as the 14 Bravais lattices. They form the
minerals. A working knowledge of certain concepts of crystallography and geometrical basis for all possible unit cells. Each Bravais lattice belongs to 1
x-ray diffraction is necessary for understanding clay mineralogy. A brief of 6 crystal systems depending on the symmetry of the lattice (Table 1-5).
introduction to some of these concepts follows. The cubic system is most symmetric. All axial translations are equal (a = b
= c) as are all axial angles (α = β = γ). The triclinic system is most general
A. Periodicity in Crystals

In a crystal, a particular pattern or arrangement of atoms is repeated over 3


The terms lattice and structure are often misused by authors and speakers. The two
and over in three dimensions. Repeating patterns occur not only in crystals terms are not synonymous. Lattice is the mathematical concept of an infinite uniform
but also in many familiar places such as wallpaper, tile floors, and brick distribution of points in space (e.g., the 14 Bravais lattices). Structure refers to the actual
physical assemblage of atoms (e.g., the 2:1 structure) (AIPEA Nomenclature Committee,
walls. Concepts that apply to crystals can be developed by starting with 1980). The concept of a lattice is used to describe the periodicity in crystal structures.
these familiar examples.
Fig. 1-10a illustrates a pattern made up of leaves much like the pattern
on a piece of wallpaper. The periodicity of this pattern can be characterized
Introductory Concepts Final Draft, Revised 1/18/00 Page 12

(and least symmetric). All axial translations and axial angles are different. point in its cycle relative to the others. The maxima of the three waves no
The unit cell dimensions and the axial angles are important parameters in longer coincide and thus the sum, which takes into account the positive and
describing the crystal structure of a mineral. negative aspects of the three individual waves, approaches zero.
In x-ray diffraction, constructive interference occurs when many waves
C. Miller Indices are in phase. The energy from the large summation wave is detected by the
diffractometer. Destructive interference occurs when the waves are out of
A consistent and concise notation is used when planes of atoms in a phase. In this case no appreciable energy is detected. The peaks of an x-ray
crystal are discussed. Different planes are generally referred to by their diffraction pattern correspond to areas where constructive interference
Miller indices, designated as (hkl). For example, we may speak of (001) occurs. Bragg's Law relates the position of these peaks to the distances
planes, (110) planes and (431) planes. The concept of Miller indices follows between planes of atoms in the crystal. This information, along with the
from the lattice concept just discussed. Consider the lattice of points in Fig. intensities of the peaks is then used for identifying mineral phases and
1-12a and a plane that is parallel to c and passes through any two lattice determining crystal structures. The following derivation of Bragg's law will
points (represented in an edge view by the heavy line). An identical plane show how it works.
must pass through each and every lattice point (light lines) because the Consider the parallel planes of atoms with indices (hkl) separated by a
lattice is periodic and each point is the same as any other. This family of distance d depicted in Fig. 1-13b. A parallel (or collimated) beam of x-rays
parallel planes cuts the a dimension of the unit cell into 2 parts and the b of wavelength λ impinges on the (hkl) planes at an angle θ. The incident x-
dimension into 3 parts. The c dimension is not cut at all (zero parts) because rays are in phase with one another. Two waves, one traveling along OA and
the planes are parallel to Z. The Miller indices are thus (230). In terms of being reflected off the upper plane of atoms and traveling along AP, and a
(hkl), the planes cut the a dimension of the unit cell into h parts, the second wave traveling along O'C and being reflected off the second plane of
b dimension into k parts, and the c dimension into l parts. Additional atoms and then traveling along CP' may, under specific conditions, be in
examples of Miller indices are given in Fig. 1-12b. phase when they reach points P and P', respectively. If the waves are in
phase at line AB, when will they be in phase at line AD and thus at PP'? A
D. X-ray Diffraction wave along CP' is in phase with a wave along AP only if the distance BC +
CD is an integral number of wavelengths. Thus
X-ray diffraction (XRD) is the main analytical technique used both to BC + CD = nλ, where n = 1,2,3,.... [1]
identify unknown mineral phases and to determine crystal structures. Soil Since BC = CD, substituting into [1] gives
minerals are usually studied using the powder diffraction method. A 2BC = nλ. [2]
powdered sample, with particles typically <50 µm in diameter, is placed in a From Fig. 1-13b,
diffractometer and irradiated with x-rays. A plot of diffracted x-ray intensity sin θ = BC/d, [3]
versus the diffraction angle — i.e. 2θ — is obtained. A simple mathematical which can be rewritten as
relationship called Bragg's Law is then used to relate the peaks on the x-ray BC = d sin θ. [4]
chart with the distances between the diffracting planes of atoms within the Multiplying both sides by 2 gives
crystals. 2BC = 2d sin θ. [5]
Substituting [2] into [5],
1. Bragg's Law nλ = 2d sin θ. [6]
This expression is Bragg's Law. Rewritten as
X-rays are electromagnetic waves with wavelengths on the order of the d/n = λ/(2 sin θ), [7]
spacings between the planes of atoms in crystals. Thus, crystals diffract x- Bragg's Law relates the perpendicular distance between diffracting planes, d,
rays. Understanding the additive relationships of waves that are in phase or to the diffraction angle θ. Since the order of diffraction, n, is not always
out of phase is important to understanding x-ray diffraction. Fig. 1-13a known, n is usually set to 1 and the quantity d/n is referred to as the
illustrates three different sine waves, all of which begin at the same point in "d-value" of a diffraction line.
their cycles. The waves are in phase because the maxima of each wave Note that the quantity needed in eq. [7] is in units of θ, while the data
coincide. The sum of all three waves is a much larger wave with the same from x-ray diffractometers is in units of 2θ. This is a consequence of the
period (distance between successive maxima) as the original waves but with physical construction of the diffractometer and division by 2 is necessary to
three times the intensity. The three waves on the right are identical to those obtain θ from 2θ data before using eq. [7] to calculate d-values. The d-
on the left, except that they are out of phase. Each wave begins at a different values can be easily calculated with a pocket calculator or a computer
Introductory Concepts Final Draft, Revised 1/18/00 Page 13

program. Tables for determining d-values from 2θ data are also given in 642, 687-704) and an understanding of these assumptions is essential. In
Brindley and Brown (1980) for several common x-ray wavelengths. addition, Lhkl measures the size of the coherently diffracting domains within
The distances between the diffracting planes in the crystal, d in eq. [7], the crystals. If the crystals consist of several coherently diffracting domains,
are fixed by the structure of the crystal. However, x-ray wavelengths (λ's) then Lhkl will be smaller than the actual physical size of the crystals observed
used for diffraction experiments are not always the same. If λ in eq. [7] by electron microscopy (Schulze and Schwertmann, 1984).
changes, θ must change because d/n is a constant fixed by the crystal
structure of the sample. X-ray line positions expressed in 2θ can only be V. SUMMARY
compared when the x-ray wavelength is the same. Calculating the d-values
for the diffraction lines allows comparison of d-values regardless of the This chapter has developed some of the terminology used later in this
wavelength. book. It has shown how the abundance of aluminosilicate minerals is a
The d-values are directly related to the unit cell dimensions and the consequence of the abundance of Si, Al, and O in the earth's crust. It has
Miller indices. The d-value of a line with Miller indices (hkl), dhkl, in the introduced the chemical and structural classification of minerals, mentioned
orthorhombic system is given by the relationship: the major minerals that will be encountered in soils, and mentioned some of
dhkl = [(h2/a2) + (k2/b2) + (l2/c2)]-1/2, [8] the environmental implications of some of the mineral. It has developed the
where a, b, and c are the unit cell dimensions. Relationships for the other major structural themes of the phyllosilicate minerals, and it has explained
crystal systems are given elsewhere (for example, Klug and Alexander, some of the crystallographic and x-ray diffraction terminology used in later
1974). chapters.
The d/n values obtained from Bragg's Law can, in most cases, be related The chapters that follow will discuss in more detail the many facets of
to actual distances between planes of atoms in the crystals. Thus, the d- soil mineralogy, particularly as it relates to current environmental concerns.
values between successive phyllosilicate layers listed in Fig. 1-7 show up in
x-ray diffraction patterns of soil clays and are used to identify the specific VI. SUGGESTED READING
clay minerals present in unknown samples. In some cases however, in
particular for randomly interstratified minerals or cases where diffraction For a more information on general mineralogy, see introductory texts
lines are very broad due to the small size of the crystals, d/n values from such as Klein and Hurlburt (1993) or Nesse (1999). Azároff (1968) and
Bragg's Law are not directly related to the distances between planes of atoms Moore and Reynolds (1997) provide good, readable introductions to
in the crystal. To explain these effects more inclusive diffraction theory crystallography and x-ray diffraction theory.
utilizing the concepts of the reciprocal lattice, structure factor, atomic
scattering factors, etc. must be applied. VII. QUESTIONS AND EXERCISES

2. The Scherrer Equation 1. Draw sketches from memory of an octahedron and a tetrahedron using
the sphere-packing, ball-and-stick, and polyhedral representations.
Once a mineral phase has been identified based on its diffraction peaks, 2. Draw sketches from memory of sphere-packing models of a tetrahedral
additional information can often be obtained from diffraction line widths. sheet, a dioctahedral sheet, and a trioctahedral sheet. (A coin makes a
Well-crystallized minerals of sand and silt size give sharp diffraction lines good template for a circle.)
whose widths are determined only by broadening caused by the x-ray 3. Draw three sketches of a unit cell and label the cell dimensions a, b, and
diffractometer itself. Clay size particles show broader lines caused by c as shown in Fig. 1-12b. On the first cell sketch in the planes with
diffraction effects from the small particles. The smaller the particles are, the Miller indices (002), on the second cell sketch planes with indices (111),
broader the diffraction lines. X-ray line width and particle size are related by and on the third cell sketch planes with indices (021).
the Scherrer equation: 4. a. Calculate the d-values of x-ray diffraction lines that occur at 8.84,
Lhkl = (Kλ)/(β cos θ), [9] 12.36, and 26.68° 2θ for a sample x-rayed using CuKα radiation (λ =
where Lhkl is the mean crystallite dimension perpendicular to the diffracting 0.15418 nm). b. Calculate the 2θ positions for these same three lines
planes with Miller indices (hkl), K is a constant equal to 0.9 if the particles when the sample is x-rayed using CoKα radiation (λ = 0.179026 nm).
are cubes, λ is the x-ray wavelength, β is the width (in radians) at one half Answers: a. 1.000, 0.716, and 0.334 nm, respectively. b. 10.27, 14.36,
peak height corrected for instrumental broadening, and θ is the diffraction and 31.09° 2θ, respectively.
line position. The application of the Scherrer equation involves a 5. A goethite sample was x-rayed with CoKα radiation. The 111
considerable number of assumptions (Klug and Alexander, 1974; p. 634- diffraction line was centered at 24.70° 2θ and its width at half-height
Introductory Concepts Final Draft, Revised 1/18/00 Page 14

(corrected for instrumental broadening) was found to be 0.35° 2θ. Use Schulze, D. G. and U. Schwertmann. 1984. The influence of aluminium
the Scherrer equation to calculate the size of the crystals (more exactly, on iron oxides: X. Properties of Al-substituted goethites. Clay Min. 19:521-
the size of the coherently diffracting domains) perpendicular to the 539.
(110) planes. Assume that K = 0.9. Answer: 27 nm. Sposito, G. 1989. The chemistry of soils. Oxford University Press, New
York. 277 p.
VIII. ACKNOWLEDGMENTS
X. LIST OF FIGURES
I thank the many students over the years that have provided feedback on
this chapter. Fig. 1-1. The structure of halite (NaCl).

IX. REFERENCES Fig. 1-2. Hexagonal closest packing of spheres in a plane.


AIPEA Nomenclature committee. 1980. Summary of recommendations Fig. 1-3. Octahedra and tetrahedra as a consequence of two planes of close-
of AIPEA nomenclature committee. Clays Clay Min. 28:73-78.
packed spheres and three ways of representing octahedra and tetrahedra.
Allen, B. L. and D. S. Fanning. 1983. Composition and soil genesis.
p. 141-192. In L. P. Wilding, N. E. Smeck, and G. F. Hall (ed.) Pedogenesis
and soil taxonomy I. Concepts and interactions. Elsevier, New York. Fig. 1-4a. A trioctahedral sheet.
Azároff, L. V. 1968. Elements of x-ray crystallography. McGraw-Hill
Book Co., New York. 610 p. Fig. 1-4b. A dioctahedral sheet.
Brindley, G. W. and G. Brown (eds). 1980. Crystal structures of clay
minerals and their x-ray identification. Mineralogical Society, London. Fig. 1-5. The tetrahedral sheet.
495 p.
Farmer, V. C., M. J. Adams, A. R. Fraser, and F. Palmieri. 1983. Fig. 1-6. (a) Oblique view of the 1:1 layer structure illustrating the
Synthetic imogolite: Properties, synthesis, and possible applications. Clay relationship between the tetrahedral and octahedral sheets. Note that
Min. 18:459-472. adjacent apical oxygens of the tetrahedral sheet (arrows ‘a’) also define
Grim, R. E. 1968. Clay mineralogy. 2nd Ed. McGraw-Hill Book Co., edges of octahedra in the octahedral sheet. Arrows marked “b” point to OHs
New York. 596 p. that lie directly in the center of the hexagonal rings of tetrahedra, although
Jackson, M. L. 1964. Chemical composition of soils. p. 71-141. In F. E. they appear off-center in this oblique view. (b) Edge view of the 1:1 and 2:1
Bear (ed.) Chemistry of the soil. Reinhold Publishing Co., New York.
layer structures illustrating phyllosilicate nomenclature. (c) An alternate
Kämpf, N., A. C. Scheinost, and D. G. Schulze. 1999. Oxide minerals.
p. F125 – F168. In: M. E. Sumner (ed.), Handbook of Soil Science. CRC edge view that arises when the 2:1 structure above is rotated normal to the
Press, Boca Raton, FL. plane of the layer.
Klein, C. and C. S. Hurlbut, Jr. 1993. Manual of mineralogy (after
James D. Dana). 21st Ed. John Wiley & Sons, New York. 681 p. Fig. 1-7. Structural scheme of soil minerals based on octahedral and
Klug, H. P. and L. E. Alexander. 1974. X-ray diffraction procedures for tetrahedral sheets.
polycrystalline and amorphous materials. 2nd Ed. John Wiley & Sons, New
York. 966 p. Fig. 1-8. Structural models of representatives of three other aluminosilicate
Marshall, C. E. 1949. The colloid chemistry of the silicate minerals. mineral groups that occur frequently in soils. All three are drawn to the same
Academic Press Inc., New York. 195 p. scale.
Moore, D. M. and R. C. Reynolds, Jr. 1997. X-ray diffraction and
identification and analysis of clay minerals. 2nd ed. Oxford University Press, Fig. 1-9. Structural details of phyllosilicates as illustrated by the octahedral
New York. 378 p. and tetrahedral sheets of muscovite. Figures were prepared from the single-
Nesse, William D. 1999. Introduction to mineralogy. Oxford University crystal structural refinement data of Rothbauer (1971).
Press, New York. 464 p.
Rothbauer, R. 1971. Untersuchung eines 2M1-Muskovits mit
Neutronenstrahlen. N. Jahrbuch f. Mineralogie. Monatsheite 1971:143-154. Fig. 1-10. (a) A lattice array of leaves and (b) a plane lattice of points.
Introductory Concepts Final Draft, Revised 1/18/00 Page 15

Fig. 1-11. A space lattice with several alternative unit cells outlined (after Ti 4.4 0.0019 0.061 (6) 0.03
Klug and Alexander, 1974). H 1.4 0.0289 ‡ ‡
P 1.0 0.0007 0.017 (4) 0.00
Fig. 1-12. (a) Planes with Miller indices 230 (c axis is out of the plane of the Mn 0.9 0.0003 0.083 (6) 0.01
figure) and (b) planes with different Miller indices on an orthogonal lattice.
† Numbers in parentheses refer to coordination number. Radii for Fe and Mn are
Fig. 1-13. (a) In phase and out of phase waves, and (b) geometry of the for the reduced (2+) form.
Bragg "reflection" analogy for diffraction of x-rays by crystals (after ‡ Ionic radius and volume of H+ are negligible compared to O2-.
Azároff, 1968).

XI. LIST OF TABLES

Table 1-1. The twelve most common chemical elements in the earth's crust
(Klein and Hurlbut, 1993).

Table 1-2. Classification of common non-silicate minerals in soils.

Table 1-3. Classification of primary silicate minerals.

Table 1-4. Ionic radius, radius ratio and coordination number of common
elements in silicate minerals (Klein and Hurlbut, 1993).

Table 1-5. Axial ratios and angles between crystal axes for the six crystal
symmetry systems.

Table 1-1. The twelve most common chemical elements in the earth's crust
(after Klein and Hurlbut, 1993).

Element Crustal Mole fraction Ionic radius† Volume


average
g kg-1 (nm) %
O 466.0 0.6057 0.136 (3) 92.88
Si 277.2 0.2052 0.026 (4) 0.22
Al 81.3 0.0626 0.039 (4) 0.23
Fe 50.0 0.0186 0.078 (6) 0.54
Ca 36.3 0.0188 0.100 (6) 1.15
Na 28.3 0.0256 0.102 (6) 1.66
K 25.9 0.0138 0.151 (8) 2.89
Mg 20.9 0.0179 0.072 (6) 0.41
Introductory Concepts Final Draft, Revised 1/18/00 Page 16

Table 1-2. Common non-silicate minerals in soils. Table 1-3. Classification of silicate minerals.

Mineral class Mineral Chemical formula† Silicate class,


Halides Halite NaCl unit composition,
Arrangement of
Sulfates Gypsum CaSO4•2H2O
SiO4 tetrahedra† Mineral Ideal Formula‡
Jarosite KFe3(SO4)2(OH)6
Sulfides Pyrite FeS2 Nesosilicates (SiO4)4- Olivine (Mg,Fe)2SiO4
Carbonates Calcite CaCO3 Forsterite Mg2SiO4
Dolomite CaMg(CO3)2 Fayalite Fe2SiO4
Nahcolite NaHCO3 Zircon ZrSiO4
Trona Na2CO3•NaHCO3•2H2O Sphene CaTiO(SiO4)
Topaz Al2SiO4(F,OH)2
Soda Na2CO3•10H2O
Garnets X3Y2(SiO4)3,
Oxides and Hydroxides X=Ca,Mg,Fe2+,Mn2+,
Aluminum Gibbsite Al(OH)3 Y=Al,Fe3+,Cr3+
Iron Hematite Fe2O3 Andalusite Al2SiO5
Goethite FeOOH Sillimanite Al2SiO5
Lepidocrocite FeOOH Kyanite Al2SiO5
Staurolite Fe2Al9O6(SiO4)4(O,OH)2
Maghemite Fe2O3
Ferrihydrite Fe5O7(OH)•4H2O Sorosilicates (Si2O7)6- Epidote Ca2(Al,Fe)Al2O(SiO4)
Magnetite Fe3O4 (Si2O7)(OH)
Manganese Birnessite (Na,Ca,Mn2+) Mn7O4·2.8
H2O Cyclosilicates (Si6O18)12- Beryl Be3Al2(Si6O18)
Lithiophorite LiAl2Mn24+Mn3+ O6(OH)6 Tourmaline (Na,Ca)(Li,Mg,Al)(Al,Fe,
Hollandite Ba(Mn4+,Mn3+)8O16 Mn)6 (BO3)3(Si6O18)(OH)4
Todorokite (Na,Ca,K)0.3-0.5(Mn4+,Mn3+)6
Inosilicates Pyroxenes
O12·3.5H2O
(single chains) (SiO3)2- Augite (Ca,Na)(Mg,Fe,Al)(Si,Al)2O6
Titanium Rutile TiO2 Enstatite MgSiO3
Anatase TiO2 Hypersthene (Mg,Fe)SiO3
Ilmenite Fe2+TiO3 Diopside CaMgSi2O6
† after Klein and Hurlburt (1993), Kämpf et al. (1999) Hedenbergite CaFeSi2O6
Pyroxenoids
Wollastonite CaSiO3
Rhodonite MnSiO3

Inosilicates Amphiboles
(double chains) (Si4O11)6- Hornblende (Ca,Na)2-3(Mg,Fe,Al)5Si6
(Si,Al)2O22(OH)2

Tremolite Ca2Mg5Si8O22(OH)2
Actinolite Ca2(Mg,Fe)5Si8O22(OH)2
Introductory Concepts Final Draft, Revised 1/18/00 Page 17

Cummingtonite (Mg,Fe)7Si8O22(OH)2 Table 1-4. Ionic radii and coordination of common elements in phyllosilicate
Grunerite Fe7Si8O22(OH)2 minerals.

Phyllosilicates (Si2O5)2- Micas


Muscovite KAl2(AlSi3O10)(OH)2 Ion Ionic radius†
Biotite K(Mg,Fe)3(AlSi3O10)(OH)2
Phlogopite KMg3(AlSi3O10)(OH)2 (nm)
Chlorites (Mg,Fe)3(Si,Al)4O10(OH)2 ..........
O2- 0.140(6)
Clay minerals
(selected) F- 0.133(6)
Talc Mg3Si4O10(OH)2
Pyrophyllite Al2Si4O10(OH)2 Cl- 0.181(6)
Kaolinite Al2Si2O5(OH)4
Smectite M+0.3Al2(Al0.3Si3.7)O10(OH)2 Si4+ 0.026(4) tetrahedral sites
M+= Ca2+, Mg2+, K+, etc.
Vermiculite M+0.7Al2(Al0.7Si3.3)O10(OH)2 Al3+ 0.054(6)
M+= Ca2+, Mg2+, K+, etc.
Serpentines Fe3+ 0.065(6)
Antigorite Mg3Si2O5(OH)4
Chrysotile Mg3Si2O5(OH)4 Mg2+ 0.072(6)
octahedral sites
0 Ti4+ 0.061(6)
Tectosilicates (SiO2) Feldspars
Orthoclase KAlSi3O8
Fe2+ 0.078(6)
Albite NaAlSi3O8
Anorthite CaAl2Si2O8
Mn2+ 0.083(6)
SiO2 Group
Quartz SiO2
Na+ 0.118(8)
Tridymite SiO2
Cristobalite SiO2
Ca2+ 0.112(8)
Zeolites
Analcime NaAlSi2O6•H2O cubic and
K+ 0.151(8)
Feldspathoids dodecahedral sites
Nephelene (Na,K)AlSiO4
Ba2+ 0.142(8)
† after Allen and Fanning (1983).
‡ Klein and Hurlbut (1993).
Rb+ 0.161(8)

† Ionic radii vary with coordination number. Radii are for ions with the
coordination numbers listed in parentheses (from Klein and Hurlbut, 1993, p.
188).
Introductory Concepts Final Draft, Revised 1/18/00 Page 18

Table 1-5. Axial ratios and angles between crystal axes for the six crystal
symmetry systems (after Klug and Alexander, 1974).

System Axial ratios Angles between crystal axes


Cubic (isometric) a=b=c α = β = γ = 90°
Hexagonal a=b„c α = β = 90°, γ = 120°
Tetragonal a=b„c α = β = γ = 90°
Orthorhombic a„b„c α = β = γ = 90°
Monoclinic a„b„c α = γ = 90°, β > 90°
Triclinic a„b„c α „ β „ γ = 90°
Inserts for Table 3

Nesosilicates
Cyclosilicates

Sorosilicates

Phyllosilicates

Inosilicates
(double chains)

Tectosilicates
Inosilicates
(single chains)
Na Cl

Fig. 1-1. The structure of halite (NaCl).


"A" voids "B" voids

Fig. 1-2. Hexagonal closest packing of spheres in a plane.


Face Edge

"A" site "B" site


Sphere-Packing
Model

Ball-and-Stick
Model

Polyhedral
Model

Tetrahedron Octahedron Tetrahedron Octahedron


Fig. 1-3. Octahedra and tetrahedra as a consequence of two planes of close-packed spheres and three ways of representing octahedra and tetrahedra.
The Octahedral Sheet (trioctahedral)

Mg2+

OH-

Figure 1-4a. A trioctahedral sheet. The upper three rows of hydroxyls in the face view of the sphere-packing model have been omitted for clarity.
The Octahedral Sheet (dioctahedral)

Al3+

OH-

Figure 1-4b. A dioctahedral sheet. The upper three rows of hydroxyls in the face view of the sphere-packing model have been omitted for clarity.
The Tetrahedral Sheet

apical oxygens
Si4+

basal oxygens

Figure 1-5. The tetrahedral sheet.


a
b
b
a

b
PLANES OF IONS SHEETS, LAYERS
basal O's
tetrahedral cations tetrahedral
sheet 1:1
OH's & apical O's
octahedral cations octahedral layer
sheet
OH's

basal O's
tetrahedral cations tetrahedral
sheet
OH's & apical O's
octahedral cations octahedral 2:1
sheet layer
OH's & apical O's
tetrahedral cations tetrahedral
sheet
basal O's

c
Alternate
edge view

Fig. 1-6. (a) Oblique view of the 1:1 layer structure illustrating the relationship between
the tetrahedral and octahedral sheets. Note that adjacent apical oxygens of the
tetrahedral sheet (arrows ‘a’) also define edges of octahedra in the octahedral sheet.
Arrows marked “b” point to OHs that lie directly in the center of the hexagonal rings
of tetrahedra, although they appear off-center in this oblique view. (b) Edge view of the
1:1 and 2:1 layer structures illustrating phyllosilicate nomenclature. (c) An alternate edge
view that arises when the 2:1 structure above is rotated normal to the plane of the layer.
Gibbsite
OH-
0.46 nm
(4.6 Å)
Al3+

Kaolinite
O2-

0.72 nm
(7.2 Å)
Si4+
Al3+

Halloysite (hydrated)

1.00 nm
H 2O
(10.0 Å)
Si4+
Al3+

Pyrophyllite
Si4+
Al3+
0.92 nm
(9.2 Å)

Mica (muscovite)
Si4+,Al3+
Al3+
1.00 nm
fixed K+
(10.0 Å)

Fig. 1-7. Structural scheme of soil minerals based on octahedral and tetrahedral sheets.
Vermiculite
Si4+,Al3+
Al3+

1.4 nm hydr. Ca2+,


(14 Å) Mg2+, etc.

Smectite

H 2O
>1.8 nm exch. Ca2+,
(>18 Å) Mg2+, etc.

Hydroxy-interlayered
Vermiculite and Smectite

Al3+ poly.,
1.4 nm exch. Ca2+,
(14 Å) Mg2+, etc.

Chlorite

1.4 nm Mg2+, Al3+,


(14 Å) Fe3+, etc.

Fig. 1-7. Continued.


Palygorskite
Si4+
Mg2+,Al3+
H 2O
H2 O
(not shown)

Clinoptilolite (a zeolite)
Si4+,Al3+

exch. cations
& H2O
(not shown)

Imogolite

Al3+
Si4+

H2O
(not shown)
1.0 nm
(10 Å)

Fig. 1-8. Sturctural models of representatives of three other aluminosilicate mineral


groups that occur frequently in soils. All three are drawn to the same scale as Fig. 1-9.
Structural Details Common to Phyllosilicate Minerals
Tetrahedral Sheet Octahedral Sheet

ditrigonal
hole
occupied
octahedra
smaller than
unoccupied

shared edges
shorter than
unshared

corrugated
basal plane

Fig. 1-9. Structural details of phyllosilicates as illustrated by the octahedral and tetrahedral sheets of muscovite. Figures prepared from the single-
crystal structural refinement data of Rothbauer (1971).
a b

g
b

a
Fig. 1-10. (a) A lattice array of leaves and (b) a plane lattice of points.
Fig. 1-11. A space lattice with several alternative unit cells outlined (after Klug and
Alexander, 1974).
Fig. 1-12. (a) Planes with Miller indices 230 (c axis is out of the plane of the figure) and
(b) planes with different Miller indices on an orthogonal lattice.

Fig. 1-13. (a) In phase and out of phase waves, and (b) geometry of the Bragg "reflection"
analogy for diffraction of x-rays by crystals (after Azároff, 1968).

You might also like