Vous êtes sur la page 1sur 30

Published 2002

Chapter 2

Surface Chemistry of Soil Minerals


CLIFF T. JOHNSTON, Purdue University, West Lafayette, Indiana

ETELKA TOMBÁCZ, University of Szeged, Szeged, Hungary

The focus of this chapter is to provide an overview of the surface chemistry of soil
mineral surfaces from three different perspectives. First, we will consider what the
reactive surface features of soil minerals “look” like on a molecular scale. The con-
cepts presented in this section are a direct extension of the structural concepts pre-
sented in Chapter 1 (Schulze, 2002). Attention here will be focused on the surfaces
of representative soil minerals. Second, we will consider how these types of sur-
face sites influence the dominant chemical and physical processes that occur in soils.
Several specific components of the soil solution, including water, metal cations, and
oxyanions, will be considered on the basis of their surface chemistry. Third, em-
phasis is placed on the surface chemistry of naturally occurring soil surfaces where
multiple solid phases, including inorganic and organic components, are present. Re-
cent attention has focused on the fact that the surface properties of soil minerals are
quite distinct from their pure mineral specimen counterparts (Zachara et al., 1993).
Current experimental evidence indicates that mineral surfaces weathered by ex-
posure to the soil solution have surface properties that are distinct from those of pure
reference minerals because of the multitude of chemical, biological, and physical
processes occurring in soils (Davis & Kent, 1990).
The solid–water interface is the dominant interface of interest to soil chemists
(Sposito, 1984; Wolt, 1994). This interface is important because it controls the re-
tention, transformation, and transport of nutrients and pollutants; provides physi-
cal support for plants and microorganisms; influences water quality; and controls
the hydraulic conductivity of soils. A conceptual illustration of the interactions at
the solid–water interface is shown in Plate 2–1. The dominant, and often neglected,
component in the soil solution is water, which is present at a concentration of about
55 M (mol dm-3). The interaction of water with soil minerals plays a critical role
in defining the physical and chemical properties of soils. Virtually all of the chem-
ical and biological reactions that occur in soil take place at the solid–water inter-
face.
Also contained in the soil solution are low concentrations of organic and in-
organic solutes (Plate 2–1). Examples of solutes commonly found in the soil solu-
tion include the dissolution products of soil minerals, metals, organic pollutants,
plant nutrients, and agricultural chemicals such as herbicides. These solutes may

Copyright 2002 © Soil Science Society of America, 677 South Segoe Road, Madison, WI 53711, USA.
Soil Mineralogy with Environmental Applications. SSSA Book Series, no. 7.

37
38 JOHNSTON & TOMBÁCZ

be: (i) positively charged, as in the case of exchangeable cations, protonated organic
bases, cationic herbicides, and certain heavy metals; (ii) negatively charged, such
as phosphate, borate, arsenate, selenate, nitrate, and ionized organic acids; or (iii)
neutral compounds, such as many weakly polar and nonpolar organic pollutants.
Depending on their size and polarity, the fate and transport of these solutes will be
influenced strongly by interactions at the solid–water interface. In addition, many
of the chemical species in the soil solution, especially those with charge, interact
with each other, adding to the complexity of soil solution chemistry. The complex,
heterogeneous nature of soil particles cannot be adequately illustrated in this sim-
ple diagram. The key points of the soil–solution interaction web are that solute be-
havior in soils is often controlled by multiple solid phases and that solute species
are influenced by the composition of the soil solution in contact with the soil par-
ticles (Buffle, 1988, p. 163–194; Benjamin & Leckie, 1981).
Depending on the nature of the solute–surface interactions shown in Plate 2–1,
the solute may undergo abiotic transformation, be strongly or weakly bound to the
solid surface, or be excluded from the solid–water interface. In addition, dissolved
organic matter, chelating agents, surfactants, or nonaqueous phase liquids in solu-
tion may compete with the soil particles for solutes, increasing the concentration
of the solute in the soil solution. These types of surface reactions dictate the poten-
tial for transformation, retention, and ultimately the transport of nutrients, agricul-
tural chemicals, and contaminants to ground and surface waters. For example, mi-
crobial degradation of herbicides in soils is controlled, in part, by the activity of the
herbicide in the soil solution. If the herbicide is retained strongly by soil particles,
it will not be accessible to soil microorganisms and will have reduced bioavailability.

I. WATER—THE UNIVERSAL SOLVENT

The importance of water and its controlling influence on the Earth’s crust is
a useful starting point to understand the surface reactivity of soils. Water is the only
commonly occurring inorganic liquid on Earth and is directly responsible for shap-
ing its surface. A similar transformation occurs on the molecular scale at the soil
particle–water interface, where the surface features of soil minerals are continually
being altered under the influence of water. Because of the asymmetric, dipolar char-
acter of water (Plate 2–2a), water molecules interact strongly with each other, with
ions in the soil solution, and with soil particles. In solution, the H atoms from one
water molecule are attracted to the O atoms of adjacent water molecules, resulting
in a web of H bonds, as shown in Plate 2–2b. As a result of these attractive forces
between water molecules, water has a remarkably high surface tension (72 mN m-1),
a wide temperature range over which it remains in the liquid state (100°C), and
unique solvation properties compared with other liquids.
In soils, water has the ability to dissolve a small amount of everything it comes
in contact with. Because it is such a powerful solvent, it is the key agent responsi-
ble for the distribution of geochemical materials around and through the Earth’s sur-
face. Of particular geochemical significance is the inclusion of atmospheric CO2
into natural waters in the form of carbonic acid. Although a very weak acid, the long-
term exposure of minerals in soil environments to this weakly acidic aqueous so-
SURFACE CHEMISTRY OF SOIL MINERALS 39

lution is partly responsible for the transformation of soil minerals. In addition, the
products of mineral dissolution reactions, such as Al3+ and Fe3+, in contact with water
result in hydrolysis and hydration reactions that are important sources of acidity.
Of equal importance is the fact that water is the prerequisite solvent for all of the
biological reactions occurring in soils.

II. ACTIVE SITES ON SOIL PARTICLES

The reactive features on soil particles can be broadly grouped into a set of
five different types of active sites. Analogous to the concept of active sites used to
describe the structure and function of biological macromolecules, such as proteins
and enzymes, this framework can be applied to better understand the surface chem-
istry of soil constituents (Sposito, 1984; Johnston, 1996; Yamagishi, 1993). These
sites encompass the dominant reactive features found on a wide range of soil par-
ticles and control most of the chemical and physical processes that occur in soils.
The term active site implies a process wherein a surface chemical reaction of in-
terest is promoted by a molecular-scale feature on the surface of a soil particle. Ac-
tive sites are described on the basis of their location (edge vs. basal surface), geo-
metric arrangement of surface atoms, chemical composition, and accessibility. A
distinction is made in this chapter between active sites on inorganic soil particles
and on soil organic matter in order to emphasize the relationship between structure
and reactivity. However, such a distinction is not made in soils where interactions
between solutes and soil particles almost always involve both inorganic and organic
surfaces.
Emphasis in Chapter 1 was on the basic structural concepts and mineralogy
of soil minerals (Schulze, 2002). In this chapter, we extend these concepts to de-
velop an understanding of the structure and reactivity of soil minerals and how these
reactive surface features interact with components of the soil solutions. This is il-
lustrated in Plate 2–3 where the crystal structure of kaolinite is illustrated in the top
portion in ball and stick (Plate 2–3a) and polyhedral representation (Plate 2–3b) sim-
ilar to the structure of kaolinite presented in Chapter 1. To illustrate the surface struc-
ture of kaolinite, a section consisting of 3 × 5 unit cells, containing 510 atoms, is
shown in polyhedral (Plate 2–3c) and space-filled (Plate 2–3d) representations. Fi-
nally, the surface charge of kaolinite is represented by an electrostatic surface po-
tential map shown in Plate 2–3e. The green colors shown in Plate 2–3e correspond
to near neutrality on the surface of the kaolinite particle. The red color shown on
the edges represents areas of local charge. These edge-sites are thought to be the
most active in terms of interacting with charged species in the soil solution. This
view is intended to show the surface charge and microtopography that a solute con-
tained in the soil solution “sees” as it approaches a particle surface.
Reactive sites on soil particles are divided into two broad categories of sur-
face sites, either polar or nonpolar sites. Polar sites may originate from isomorphic
substitution sites and edge sites of minerals, or ionized functional groups of soil or-
ganic matter. Nonpolar sites include interactions on the surfaces of neutral mineral
surfaces and with the nonpolar, or hydrophobic, portion of soil organic matter. Table
2–1 shows the qualitative distribution of active site types on common soil con-
40

Table 2–1. Relative distribution of active sites on selected soil components.


Polar sites
Permanent charged Conditionally charged Nonpolar sites
Isomorphic Broken edges Ionizable organic Neutral siloxane Humic organo-
Solid phase substitution terminal OH groups functional groups surface phyllic phase
Organic matter high high
Carbonates low
Aluminum oxides and hydroxides high
Allophane and imogolite high
Iron oxides high
Manganese oxides high
Kaolin and serpentine minerals low intermediate intermediate
Phyrophyllite and talc low high
Mica and illite low low low
Smectite high intermediate low intermediate low
Vermiculite very high intermediate low
Chlorites intermediate low
Palygorskite and sepiolite high
Zeolites low to high low to high low to high
Silica minerals high intermediate
Titanium and zirconium minerals low
JOHNSTON & TOMBÁCZ
SURFACE CHEMISTRY OF SOIL MINERALS 41

stituents. In the following section, active sites of soil particle are presented in two
groups. Inorganic surface sites are introduced first, beginning with the neutral
siloxane surface, followed by the constant charge sites resulting from isomorphic
substitution, and finally the reactive inorganic surface hydroxyl group. In the sec-
ond group, the active sites of soil organic matter, both polar and nonpolar sites, are
introduced. One of the objectives of this section is to associate the type and distri-
bution of active sites with various types of soil minerals.

A. Active Sites on Inorganic Particles

1. Neutral Siloxane Surface


From a surface reactivity standpoint, the least reactive type of inorganic sur-
face occurring in soils is the neutral siloxane surface. These surfaces have no
charge and no permanent dipole moment. As a result, this surface does not inter-
act strongly with the H bond network of water molecules and is termed a hy-
drophobic surface. The literal definition of hydrophobic is “water fearing”, and refers
to a surface that is water repellant. At the opposite end of the spectrum, a hydrophilic
surface represents a “water loving” surface. The beading of water on an oily or
waxed surface, for example, is the macroscopic manifestation of water interacting
with a hydrophobic surface. On a molecular scale, water molecules are repelled,
or pushed away, from this type of surface. Because water molecules have little, or
no, interaction with a hydrophobic surface, a hydrophobic surface is not able to form
H bonds with H2O. Hydrophobic surfaces occur on 2:1 phyllosilicates where no iso-
morphic substitution has occurred, such as in talc and pyrophyllite, and on the silox-
ane-side of kaolinite. They consist mainly of Si–O bonds where the Si is in tetra-
hedral coordination. In this configuration, the -2 charge of each O atom is
completely satisfied by the Si atoms bonded to it (Huheey, 1978).
The siloxane surface of neutral phyllosilicates functions as a very weak
Lewis base. A surface Lewis base is defined as a feature that can furnish a pair of
electrons for a chemical bond (Jensen, 1978). However, the electron-donating abil-
ity of the neutral siloxane surface is quite limited. Neutral siloxane surfaces have
a low affinity for water. One way to measure the affinity of a liquid with a surface
is to measure the contact angle. When a liquid is placed on a solid surface, the ten-
dency of a drop to spread or bead up on the surface is reflected by the angle of con-
tact between the liquid and the solid phase (Adamson, 1989) (Fig. 2–1). For well-

Fig. 2–1. Illustration of the solid–liquid contact angle showing a drop of liquid on a solid surface.
42 JOHNSTON & TOMBÁCZ

defined, flat surfaces, contact angle measurements can be used to determine the hy-
drophobic–hydrophilic nature of a surface. For liquids that are attracted to the sur-
face, the contact angle is small, with values <20° corresponding to a hydrophilic
surface. For example, the contact angle of water on vermiculite ranges from 0 to
15° (Yariv, 1992), clearly in the hydrophilic range. In contrast, the contact angle of
water on talc and pyrophyllite is >80°, indicating the surface is hydrophobic. Since
most of the surfaces that occur in soils are not flat, or well-defined, contact angle
measurements have seen little application in soils. However, the hydrophobic–hy-
drophilic nature of soil particles is important and is related to such applied prob-
lems as the water repellant soils that can develop under citrus (Citrus spp.) and eu-
calyptus (Eucalyptus spp.) trees.
From a surface reactivity standpoint, the neutral siloxane surface is inert.
Solute interactions with this type of surface are restricted to van der Waals inter-
actions, weak, nonspecific surface interactions resulting from a combination of fac-
tors, such as dipole–dipole and dipole–polarization interactions, also called Lon-
don forces. Although polar molecules, including water, and aqueous electrolytes
have a low affinity for the neutral siloxane surface, nonpolar organic solutes and
the nonpolar portion of larger biological molecules such as proteins and enzymes
can efficiently bond to this type of surface through van der Waals forces.
On a more restricted spatial scale, the neutral siloxane surface also occurs on
charged 2:1 phyllosilicates in the region between isomorphic substitution sites. Re-
cent pesticide sorption studies (Chapter 26, Laird & Sawhney, 2002) have shown
this type of surface on charged phyllosilicates can significantly stabilize organic
solutes that have both a polar and nonpolar character. Molecules such as atrazine
[6-chloro-N-ethyl-N¢-(1-methylethyl)-1,3,5-triazine-2,4-diamine], for example,
have both polar and nonpolar regions and can be strongly sorbed to low charge ex-
pandable clay minerals. The neutral siloxane surface stabilizes the nonpolar por-
tion of the molecule, while the polar portion of the molecule interacts directly with
the isomorphic substitution site or with waters of hydration surrounding the ex-
changeable cation.
2. Constant Charge Sites
Permanent charge sites result from isomorphic cationic substitutions in the
crystal structure of 2:1 phyllosilicates (Plates 2–4 and 2–5). These sites are char-
acterized by a permanent negative charge and hence are referred to as constant
charge sites. They occur on the basal surfaces of charged phyllosilicates, includ-
ing smectites, vermiculites, chlorites, and micas. Isomorphic substitution can occur
in either the octahedral or tetrahedral sheet of 2:1 phyllosilicates, and many clay
minerals are characterized by varying degrees of substitution in both sheets. In the
case of Mg2+ for Al3+ substitution in the octahedral sheet (top portion of Plate 2–4),
the charge is delocalized over a relatively large region thought to be about nine O
atoms on the siloxane surface, according to Sposito (1984). These two types of iso-
morphic substitution sites are illustrated by the dashed (tetrahedral site) and dot-
ted circles (octahedral site) shown in Plate 2–5. Substitution in the tetrahedral sheet,
commonly Al3+ for Si4+, results in a more localized charge distribution over ap-
proximately three oxygens, illustrated by the single thick negative signs in Plate
2–4 and by the dashed circle in Plate 2–5.
SURFACE CHEMISTRY OF SOIL MINERALS 43

Plate 2–1. Conceptual interaction web showing a soil particle comprised of clay minerals, oxides, and
humic components interacting with the soil solution.

Plate 2–2. Illustration of (A) the charge on a water molecule illustrating the dipolar nature and (B) the
structure of water molecules in solution showing the network of H bonds.
44 JOHNSTON & TOMBACZ

Plate 2–3. Representation of the kaolinite surface shown in (A) ball-and-stick and (B) polyhedral,(C)
a 5 by 5 by 0.7 nm section of the kaolinite surface in polyhedral, and (D) ball-and-stick and space-
filled models. (E) is an electrostatic surface potential map of the kaolinite surface (courtesy of Randy
Cygan at Sandia National Laboratory).

Plate 2–4. Edge view of a 2:1 phyllosilicate showing isomorphic substitution in the octahedral sheet (top)
and tetrahedral sheets (bottom).
SURFACE CHEMISTRY OF SOIL MINERALS 45

Plate 2–5. A 1.6 by 1.6 nm portion of the siloxane surface is shown on the left side in ball-and-stick
representation. The total electron density of this structure is shown on the right side, illustrating the
size of the ditrigonal cavity.

Plate 2–6. Hydrated cation near the surface of a 2:1 phyllosilicate. Example of an outer-sphere surface
complex.
46 JOHNSTON & TOMBACZ

Plate 2–7. Edge sites of gibbsite. The ball-and-stick structure of gibbsite, Al(OH)3, is shown projected
on edge (along the crystallographic b axis) at the top. (A) A larger portion of the crystal; (B) devel-
opment of an edge-site, leaving two dangling Al–O bonds. (C) In the larger crystal, these bonds were
connected to Al atoms with each Al atom providing a charge of +0.5. These dangling bonds are now
undercoordinated and will react with water or other solutes in an attempt to complete their coordi-
nation sphere.

Plate 2–8. Three-dimensional structure of humic acid shown using a ball and stick representation on
the left side and as the electrostatic potential surface on the right side. The gray regions of this sur-
face correspond to neutral portions of humic acid, blue regions represent negatively charged areas
and red regions to portions of the humic acid surface with a slight positive charge. Using the Mole-
cular Simulations Cerius2 and Weblab ViewerPro programs, structures were plotted based on the two-
dimensional structure of Schulten and Schnitzer (1995).
SURFACE CHEMISTRY OF SOIL MINERALS 47

Plate 2–9. Example of an outer-sphere surface complex—hydrated cation near the surface of a 2:1 phyl-
losilicate.

Plate 2–10. Example of an inner-sphere complex—phosphate bound to the exposed Fe–O groups on
the surface of goethite.
SURFACE CHEMISTRY OF SOIL MINERALS 49

To better understand the chemical and physical properties of reactions in-


volving constant charge sites, an expanded view of these sites is shown in Plate 2–5.
A 1.6 by 1.6 nm section of the siloxane surface is shown in ball-and-stick repre-
sentation on the left side of Plate 2–5. The same structure and scale are shown on
the right side as the solvent accessible surface. In this representation, the surface
topography of the siloxane surface is revealed, showing the hexagonal network and
ditrigonal cavities. The solvent accessible surface is the surface an exchangeable
cation “sees”. An important feature of this surface is the size of the siloxane ditrig-
onal cavity. For comparison, the ionic radii of K+, Na+, Ca2+, and Mg2+ are shown
in the lower portion of Plate 2–5. These ditrigonal (i.e., pseudohexagonal shaped)
cavities have a diameter of about 0.23 nm (right side of Plate 2–5), which is close
to the ionic radii of both K+ and NH4+ ions. This close match in size of the K+ and
NH4+ ions to that of the siloxane ditrigonal cavity is responsible, in part, for K+ and
NH4+ fixation in soils. The siloxane surface is ubiquitous in nature and occurs on
all phyllosilicates. It is an important surface feature because the charge deficit that
occurs in smectites, vermiculite, chlorites, and micas is manifested on the siloxane
surface.
Depending on the extent of isomorphic substitution, these negatively charged
sites are separated by distances ranging from 1 to 2 nm on the basal surface (Fig.
2–2b). The negative charge is balanced by cations such as Ca2+, Mg2+, Na+, and K+
external to the 2:1 sheet, with Ca2+ being the dominant exchangeable cation oc-
curring in most soils. This is illustrated in Fig. 2–2 showing a 5 by 5 by 1 nm sec-
tion of a 2:1 phyllosilicate. The upper structure (Fig. 2–2a) depicts a neutral sur-
face where no isomorphic substitution has occurred and would be representative
of the neutral siloxane surface. In the case of smectites, vermiculites, and micas,
minerals in which isomorphic substitutions do occur (Fig. 2–2b), the correspond-

Fig. 2–2. Basal surface (001) of a 2:1 phyllosilicate: (A) that of a neutral surface such as talc or pyro-
phyllite where no isomorphic substitution has occurred; (B) the approximate location of the isomor-
phic substitution sites spread over six SiO4 tetrahedral units; (C) exchangeable cations and (D) hy-
drated exchangeable cations.
50 JOHNSTON & TOMBÁCZ

ing charge densities range from 0.1 to 0.3 C m-2. Combined with the large specific
surface areas (600–800 m2 g-1) of expandable clay minerals, constant charge sites
are among the most reactive inorganic surface sites occurring in soils on a mass basis,
with cation exchange capacities up to 150 cmolc kg-1.
For expandable phyllosilicates, such as smectites and vermiculites, the ex-
changeable cations do not occur as “bare” cations on the soil–water interface.
Rather, these cations are surrounded by a cluster of water molecules attracted to the
exchangeable cation, as shown in Fig. 2–2d. Because of their close proximity to
the metal cation, these water molecules have chemical and physical properties that
are distinct from those of bulk water. Their mobility is more restricted due to po-
larization effects by the cation and, depending on the nature of the interlayer cation,
can be more acidic than that of bulk water. The hydration characteristics (i.e., po-
larizing power) of the exchangeable cations influence many of the important chem-
ical and physical properties of soils, including the ability of soil minerals to shrink
when dry and swell when wet, retention of moisture, retention and release of nu-
trients to plants, and the formation of stable soil aggregates.
a. Influence of Exchangeable Cations on Permanent Charge Sites. The
presence of hydrated exchangeable cations in the interlayer of expandable 2:1
phyllosilicates imparts an overall hydrophilic nature to the clay surface. All inor-
ganic cations attract water molecules because of their charge, resulting in the for-
mation of water clusters around the cations. This process is termed hydration, and
the hydration energies of common cations are listed in Table 2–2 as enthalpies of
hydration, the energy released when a metal ion in the gas phase is placed in water.
As the hydration energy of the cation increases, the strength of the interaction be-
tween the cation and the water molecules surrounding the cation is increased.
Thus, water molecules are more easily removed from Cs+ than from Fe3+. Plate 2–6
illustrates hydrated exchangeable cations near a clay surface. Depending on the hy-
dration energy and hydrolysis constant of the exchangeable cation, a varying num-
Table 2–2. Hydration energies of common cations.
Effective ionic Ionic potential‡ Enthalpy of Hydrolysis
Ion radius† (reff) Z2/reff hydration§ constant¶
nm kJ mol-1 pKh
Ca2+ 0.126 2.15 -1669 12.7
Mg2+ 0.086 2.8 -1998 11.4
K+ 0.165 0.45 -360
Na+ 0.132 0.5 -444 14.5
Cs+ 0.188 0.4 -315
Cu2+ 0.087 2.6 -2174 7.5
Fe3+ 0.069 6.3 -4491 2.2
Fe2+ 0.075 2.5 -2009
Co2+ 0.079 2.6 -2106 9.6
Zn2+ 0.088 2.6 -2106 9.6
Cd2+ 0.124 2.3 -1882 11.7
Pb2+ 0.143 2 -1556 7.8
Al3+ 0.067 6.9 -4774 5.1
Li+ 0.090 0.7 -559 13.8
† Shannon (1976). § Friedman and Krishnan (1973).
‡ Huheey (1974). ¶ Yatsimirksii and Vasil’ev (1960).
SURFACE CHEMISTRY OF SOIL MINERALS 51

ber of water molecules will be strongly associated with the cation. For cations with
a large hydration energy, a larger number of water molecules (~6–12) will be at-
tracted to the cation, resulting in an overall increase in the effective radius of the
cation. For cations with lower hydration energies (Cs+, for example), less energy
is required to displace water molecules from the cation, allowing the cation to ap-
proach the surface more closely. This results in a stronger interaction of the cation
with the surface. As shown in Table 2–2, the exchangeable cations commonly found
in soils, Ca2+, Mg2+, K+, and Na+, represent a wide range of hydration energies. The
chemical nature of these exchangeable cations (i.e., effective ionic radius, hydra-
tion energy, hydrolysis constant) determines many of the important chemical and
physical properties of soils.
Hydrolysis reactions occur when the O–H bond of water molecules coordi-
nated to a metal cation are ruptured with ionization of the hydrate to yield hydro-
nium ions (i.e., H3O+). Cations that readily undergo hydrolysis reactions are small
in size and/or have a high charge. The aluminum cation, Al+3(aq), has both a small
size and a high charge (+3) and readily hydrolyzes water as shown by the reaction,

Al+3 + 6H2O ] [Al(H2O)6]+3 ] H3O+(aq) + [Al(H2O)5(OH)]+2

In contrast to Al3+, Na+ is a weakly hydrolyzing cation with a pKh of 14.5.


Recall that pKh refers to the negative log10 of the hydrolysis constant, so that a large
pKh value corresponds to a very small number. The large pKh value indicates that
Na+ has minimal ability to hydrolyze water (i.e., rupture the O–H bond). On the
other hand, Na has an intermediate ionic hydration energy of -444 kJ mol-1. These
chemical properties of the Na+ ion result in the formation of a Na+ cation surrounded
by a cluster of water molecules in the interlayer region of the clay. The hydrated
radius of the ion–water cluster prevents the penetration of the Na+ ion into the silox-
ane ditrigonal cavity (shown on the right side of Plate 2–5). For 2:1 phyllosilicates
with intermediate charge (e.g., smectites), the hydrated cation serves as a hy-
drophilic pillar, which in turn attracts additional water molecules. The amount of
water sorbed by expandable clay minerals depends strongly on the surface charge
density of the phyllosilicate and on the hydration energy of the exchangeable
cation.
b. Electrified Interface of Montmorillonite. For clays with permanent
charge sites, such as smectites, it is useful to consider the density charge sites on
the basal surface of the clay. Consider, for example, a montmorillonite with a spe-
cific surface area of 800 m2 g-1 and a cation exchange capacity of 80 cmolc kg-1.
First, we assume all of the charge originates from permanent charge isomorphic sub-
stitution sites such as those shown in Plate 2–4 and Fig. 2–2b and that all sites are
accessible for cation exchange. The cation exchange capacity is converted to the
number of charge sites per square nanometer of surface, as shown using Eq. [1].

cmol 1 kg 0.01 molcharge sites 6.02 × 1023 sites


80 ____c ______ ______________ _____________
‰ kg  ‰ 1000 g  ‰ cmolcharge sites  ‰ molcharge sites 
_____________________________________________ 0.6 sites
= ______ [1]
800
______m2 1018 nm2
1_________
× ‰ nm 2 
‰ g  ‰ m2 
52 JOHNSTON & TOMBÁCZ

Another way of looking at the site density is by taking the inverse of 0.6 sites
nm-2, which is 1.6 nm2 site-1. Assuming that each M+ ion occupies a square foot-
print of 1.6 nm2, the average distance between each M+ ion is obtained by taking
the square root of this number, which is 1.26 nm (see Fig. 2–2c).
Charge density values, so, are directly related to the site density of constant
charge sites. Charge density is expressed as coulombs per square meter (C m-2).
We can convert our site density from above to charge density using Faraday’s con-
stant as shown in Eq. [2].

cmol 1 kg 0.01 molc 96500 C


80 ____c ______ ________ _______
‰ kg  ‰ 1000 g  ‰ cmolc  ‰ molc 
_____________________________________ 0.096 C
= ______ [2]
800 m
______2 ‰ m2 
‰ g 

Charge density is sometimes represented as microcoulombs per square meter, and


the value of 0.096 C m-2 is equivalent to 9.6 μC cm-2.
In addition to the large surface areas associated with expandable clay min-
erals, such as montmorillonite, the intrinsic charge on these surfaces has a profound
influence on solute interactions in soils. A direct consequence of this charged char-
acter is the high cation exchange capacity and water retention values of these ma-
terials. The hydrated, exchangeable cations attract and hold water molecules near
the surface. In addition, the cations do not “screen” (completely neutralize) all of
the electrical charge at the surface. Thus, some of the negative charge extends into
the soil solution. This property attracts positively charged species and repels the neg-
atively charged ones. Finally, the presence of hydrated cations near the surface im-
parts a pronounced hydrophilic character to the surface. Traditionally, hydropho-
bic solutes, including most pesticides and organic pollutants, have been considered
to have minimal interaction with these types of surfaces. However, recent studies
have shown that certain types of organic compounds (e.g, nitroaromatic and tri-
azines) have appreciable sorption on low-charge smectites (Laird et al., 1992;
Haderlein & Schwarzenbach, 1993).

3. Conditionally Charged Sites—The Inorganic Surface Hydroxyl Group


As explained in Chapter 1 (Schulze, 2002), oxides of Fe, Al, Mn, Si, and Ti
are commonly found in soils. One of the important features of these soil minerals
is the development of surface charge that varies with pH. The reactive sites on these
minerals are referred to as conditionally charged sites, variable charge sites, or pH
dependent sites (Sposito, 1984). In addition, similar sites also occur on the edges
of 2:1 and 1:1 phyllosilicate clay minerals.
To illustrate how charge develops on conditionally charged sites consider the
structure of gibbsite, Al(OH)3 (Plate 2–7), a commonly occurring Al oxide in
soils. Aluminum has a valence of +3 and is in sixfold coordination to six OH groups.
Each of the OH groups in gibbsite is bonded to two Al atoms. From the standpoint
of an O atom located in the interior of a gibbsite crystal, well away from the edge,
the hydrogen attached to the O atom contributes a charge of +1 and each of the two
SURFACE CHEMISTRY OF SOIL MINERALS 53

Al bonds contributes a charge of +0.5, exactly balancing the -2 charge of the oxy-
gen. Hydroxyl groups located at the edge of the gibbsite crystal are different, how-
ever, and are called terminal OH groups. The terminal OH groups (Plate 2–7) are
bound to only one Al atom and have a charge of -0.5 because of the absence of a
second Al–O bond. These terminal OH groups are referred to as being undercoor-
dinated. Addition of an H+ atom simply results in a charge of +0.5. In other words,
the charge on the terminal OH group is unsatisfied. Because these undercoordinated
edge sites always have either a partial positive or partial negative charge, these min-
erals are more reactive on their edges than on their basal surfaces.
In addition to conditionally charged terminal OH groups, neutral inorganic
surface hydroxyl groups also occur on the basal surfaces of gibbsite and 1:1 phyl-
losilicates, as well as goethite and other oxides. The hydroxyl groups are coordi-
nated to metal atoms whose coordination environment is complete and who do not
possess a permanent charge. These groups are not chemically reactive because their
coordination environment is complete.

4. Surface Charge Density


Soil particles are characterized by two important features related to soil
chemistry. As noted above, soil particles are characterized by large surface areas
resulting from the small size of the secondary phyllosilicates and minerals in the
clay fraction. However, in addition, the interface between the soil solution and the
soil particles is characterized by an electrical charge. Thus, soil particles are char-
acterized by an electrified surface distributed over a large surface area. An impor-
tant physical property related to the development of charge on soil particles is the
concept of surface charge density (Sposito, 1984). According to Sposito, the sur-
face charge density of a soil particle can be expressed as:

sin = so + sH [3]

where sin is the intrinsic surface charge density, so is the permanent structural charge
density, and sH is the net proton surface charge. Each of these terms can be mea-
sured and can have positive or negative values. The units are typically in coulombs
per square meter, microcoulombs per square cm, or in moles of charge per square
meter. It should be noted, however, that measurement of sH is difficult and not very
accurate. The origin of the permanent structural charge density term so in soils
(which is always negative) is due to constant charge sites and these sites are always
negatively charged. Depending on the type of surface and the pH of the soil solu-
tion, however, the value of the net proton surface charge term sH can be positive
or negative (Sposito, 1984).
Inorganic surface hydroxyl groups, along with the ionizable organic functional
groups on humic substances, are responsible for the development of the pH de-
pendent charge in soils and sediments. They are amphoteric; that is to say they de-
velop a positive charge at low pH values and a negative charge at high pH. As pH
increases, these sites develop a neutral charge and ultimately a negative charge. The
overall contribution of these sites to the cation exchange capacity depends strongly
on the size, shape, and mineralogy of the particles, and the pH and ionic composi-
54 JOHNSTON & TOMBÁCZ

tion of the soil solution. With the exception of zeolites, the cation exchange capacity
originating from large particles (i.e., sand- and silt-size particles) is minimal. As
the particle size decreases, however, the contribution of variable charge sites to the
overall reactivity of the particle may become appreciable. The oxides and oxyhy-
droxides commonly found in highly weathered soils are dominated by pH depen-
dent Al–OH and Fe–OH groups and have little or no permanent charge.
Naturally occurring Fe oxides, such as ferrihydrite, goethite, and hematite,
are often among the smallest particles found in soils. Because of their small size,
the majority of the Fe–OH groups contained in these samples are exposed near the
surface. Consequently, these particles contribute significantly to the total amount
of reactive conditionally charged terminal OH groups in soils even when they com-
prise only a small percentage of the soil on a total mass basis (Schwertmann,
1990). The points of zero charge (PZC) of various Fe oxides are typically around
8 and are essentially independent of their mineralogical form. The PZC values of
soil Fe oxides, in general, are lower than that of synthetic samples, usually below
pH 7, because of specifically adsorbed anions such as silicates, phosphates, organic
anions, and carbonates from the soil solution. Iron oxide surfaces are important be-
cause of their ability to specifically adsorb anions such as phosphate and for their
interaction with other solid phases resulting in the aggregation and cementation of
soil particles.

B. Active Sites on Soil Organic Matter

1. Ionizable Organic Functional Groups of Humic Substances


Humic substances are distinctive to soil environments in that they are dis-
similar to the biopolymers of microorganisms and higher plants (including lignin)
and other small molecular weight exudation and decomposition products such as
organic acids and sugars. They are defined as a complex series of relatively high
molecular weight, yellow to black colored organic substances formed by sec-
ondary synthesis reactions in soils. Most of the organic C in soils occurs as humic
substances. The term is used in a generic sense to describe the colored material or
its fractions obtained on the basis of solubility characteristics. Humic substances
have a critical influence on the properties of soil minerals (Oades, 1989).
Most of the humic substances found in soil are chemically bound to clay min-
erals and oxides (Plate 2–1). A small fraction can be dissolved in the soil solution,
especially under alkaline conditions. Important characteristics of humic materials
are their ability to form water-soluble or water-insoluble complexes with metal ions,
hydroxides, oxides, and clay minerals, and their ability to interact with organic com-
pounds such as alkanes, fatty acids, surfactants, and pesticides.
The structure and composition of soil organic matter is discussed in detail by
Stevenson (1994). Of particular interest are the active sites present on soil organic
matter. Two broad types of active sites are considered here, (i) ionizable organic
functional groups and (ii) nonpolar van der Waals sites. Both types of sites can be
illustrated using the structural model recently developed by Schulten and Schnitzer
(1995) as a prototype of humic acid. There is still considerable controversy and dis-
agreement about the specific structure and morphology of humic substances, but
SURFACE CHEMISTRY OF SOIL MINERALS 55

certain basic structural features of soil organic matter are known. By combining the
known elemental composition of humic substances (Aiken et al., 1985) with 13C
nuclear magnetic resonance (NMR), various types of mass spectrometry, and re-
lated spectroscopic methods, Schulten and Schnitzer (1995, 1997) have developed
a representative structure for humic acid (HA) (Fig. 2–3) with the proposed formula
C308H335O90N5. One limitation of this structure is that the C/N ratio is 62:1; typi-
cal values for soil organic matter are in the range of 20 to 11:1. Nonetheless, this
two-dimensional structure can be transformed into a three-dimensional structure
as shown in Plate 2–8 using methods of computational chemistry. The three-di-
mensional structure is shown as a ball-and-stick representation on the left side and
as the solvent accessible surface on the right.
Results from the oxidative and reductive degradation of humic and fulvic acids
reveal that aromatic rings and aliphatic chains are important structural units in humic
substances. Aromatic rings are thought to be crosslinked by aliphatic chains. Liq-
uid- and solid-state 13C NMR spectra of humic substances have provided additional
evidence that aliphatic structures are important structural components of humic sub-
stances. These “regions” of the HA structure are nonpolar (Fig. 2–3) and capable
of interacting with nonpolar organic compounds. In addition, fluorescence quench-
ing studies have provided spectroscopic evidence that HAs form cage-like,

Fig. 2–3. Two-dimensional structure of humic acid from Schulten and Schnitzer (1995). The light-gray
shaded areas correspond to nonpolar regions on the molecular surface and the dark-gray shaded por-
tions to polar regions.
56 JOHNSTON & TOMBÁCZ

pseudomicellar structures along their hydrocarbon chains, even in dilute solution,


similar to that shown in Plate 2–8.
The ionizable functional groups of soil organic matter include carboxyl,
amine, phenolic, ketonic, sulfhydryl, and sulfonic. Of these, the most abundant and
reactive group is the carboxylate group, which is largely responsible for the in-
trinsically high cation exchange capacity of humic substances. Inspection of the two-
dimensional structure of HA shown in Fig. 2–3 reveals 21 carboxylate groups. The
presence of O, N, and S atoms in the structure of HA (only O and N were consid-
ered by Schulten and Schnitzer [1995]) will impart a polar character to the HA mol-
ecule. Examples of these polar regions are shown in the two-dimensional repre-
sentation of HA (Fig. 2–3) by the dark-gray shaded regions and in Plate 2–8, where
the blue region corresponds to negatively charged sites and the red to positively
charged sites. Depending on the type of C atom to which the carboxylate group is
attached to, it can be moderately to weakly acidic. Similar to the inorganic surface
hydroxyl groups found on soil minerals, these ionizable surface functional groups
make a significant contribution to the variable charge nature of soils.
The most striking feature of the solvent accessible structure in Plate 2–8 is
its three-dimensional character and large number of reactive functional groups. Be-
cause of its complex structure, however, not all of these sites are available to the
soil solution. Some of these sites will be involved in attachment of the HA mole-
cule to a mineral surface, while others will be “hidden” within the internal struc-
ture. Humic substances have a very strong influence on the overall reactivity of min-
eral surfaces. Current evidence suggests that a significant portion of exposed
mineral surface area in soil environments are impacted by the presence of humic
substances. Measurement of soil organic C in mineral soils with low specific sur-
face areas (<10 m2 g-1) can be difficult because of the small amount of C present.
An important role of humic substances in soil is to aggregate and bind inor-
ganic mineral phases together. In aqueous solution, humic acids can be dissolved
or precipitated depending on the pH and electrolyte present. In soils, these disso-
lution and precipitation processes take place continuously (Tombacz, 1990). Be-
cause of their large size and negative charge, HAs are excluded from the interlayer
regions of expandable phyllosilicates. Consequently, humic substances reside on
the external surfaces of soil minerals and strongly influence the chemical behav-
ior of solutes in the soil solution. The size, shape, and aggregation state of humic
substances are influenced by changes in water content, pH, and ionic strength.

2. Nonpolar Sites of Humic Substances


Humic substances possess a dual character. As noted above, the polar func-
tional groups of humic substances are responsible for their ability to retain metal
cations and polar compounds. However, in addition, humic substances are the
dominant phase in soils responsible for the sorption of nonpolar hydrophobic or-
ganics. How can this dual nature be attributed to the same material? Inspection of
the electrostatic potential surface of the HA structure of HA (Fig. 2–3 and Plate 2–8)
reveals the presence of both polar and nonpolar regions.
The gray-colored regions on the electrostatic potential surface of HA (Plate
2–8, left side) correspond to “neutral” or nonpolar regions on the molecular struc-
SURFACE CHEMISTRY OF SOIL MINERALS 57

ture. These regions correspond to alkyl chains and phenyl rings where O and other
polar atoms are not present. These nonpolar regions provide a hydrophobic mi-
croenvironment where nonpolar organic solutes can interact with HA molecules
through van der Waals bonding interactions. Depending on the conformation of the
HA molecule, it is possible for nonpolar organic solutes to become entrapped in the
hydrophobic interiors of HA structures.
The dual nature of humic substances (i.e., amphiphilic character where both
hydrophilic and hydrophobic parts exist) is thought to be responsible for the ionic
surfactant-like solution properties of HAs. Ionic surfactants have both a polar head
group (can be negatively or positively charged) and a hydrophobic tail that is nor-
mally an alkyl chain. In humate solutions of higher concentration, micelles can form
where clusters of HA molecules form with the hydrophobic portions of HA mole-
cules aligned towards the interior of the cluster. Water insoluble organic compounds,
such as the pesticide DDT [1,1,1-trichloro-2,2-bis(4-chlorophenyl)ethane] and the
organic pollutant benzopyrene, can be adsorbed in this hydrophobic interior region
of humate micelles. Surface coatings of humic materials are often found on the min-
eral grains in soils. A membrane–micelle humus model for organic coatings in soil
was introduced by Wershaw (1993) to describe the bilayer–lamellar arrangement
of humic substances on a mineral surface.

C. Active Sites on Natural Soil Surfaces

The surfaces of soil particles are continually being altered by natural processes
initiated by biological, biogeochemical, and hydrological agents. These processes,
known as chemical weathering, significantly enhance the surface heterogeneity of
soil particles (Newman & Hayes, 1990; Bolt & Bruggenwert, 1978; Sposito, 1984).
Specifically, the nature and reactivity of active sites on solid surfaces are altered
through processes such as the dissolution and precipitation of inorganic particles
and degradation and synthesis of organic material. Soils are characterized by open
boundaries to the atmosphere, lithosphere, and hydrosphere, and this is reflected
in the transience of soil properties (Jenne, 1980). This transience is also reflected
on the molecular scale in the distribution and type of active sites. Minerals in young
soils are generally characterized by a predominance of constant-charge, siloxane
ditrigonal cavities. Pedochemical weathering of soils alters this distribution of
sites such that the dominant reactive sites are inorganic surface hydroxyl groups
(Sposito, 1984).
Important examples of surface altering processes that occur in soils include
transformations involving Fe and Al. Specific adsorption of Fe3+, for example, oc-
curs on the surface of kaolinite at low pH, followed by the polymerization of un-
stable Fe3+ monomers (Sanchez et al., 1990). Iron-hydroxy-polycations existing at
low pHs can interact with negatively charged clay and oxide particles, resulting in
aggregation and cementation of soil particles (Schwertmann, 1990). Aluminum re-
leased to the soil solution can be adsorbed on permanent charge sites of expand-
able clay minerals and subsequently polymerized to form hydroxy Al polymers.
The presence of natural organic acids in the soil solution can complex Al ions and
hinder the formation of these hydroxy Al polymers. The formation of hydroxy
Al–montmorillonite complexes is perturbed significantly by HAs. This inhibition
58 JOHNSTON & TOMBÁCZ

effect is a possible reason for limited Al interlayering of clays in organic rich soils
(Huang, 1990).
Each of the active sites presented earlier is characterized by a distinct surface
charge that varies with pH and ionic composition of the soil solution. Because soils
are characterized by a heterogeneous distribution of active sites, two limiting cases
of heterogeneity can be considered (Koopal & vanRiemsdijk, 1989; Koopal, 1996).
If the different types of sites are mixed in a random way, the surface takes on a uni-
form charge that represents a composite of the individual components. However,
if the different classes of sites are grouped together in patches, the surface hetero-
geneity is patchwise. Examples of patchwise heterogeneity include, for example,
the distribution of Fe and Mn oxides in soils, which tend to occur as discrete con-
centrated aggregates as opposed to a smeared out distribution over the entire sur-
face of the soil particle. The size of patches and the lateral interactions are impor-
tant to describe the patchwise heterogeneity.
In addition, the diversity of active sites present on soil surfaces contributes
to the failure of simple models, like the Gouy–Chapman model, to adequately de-
scribe the chemical reactivity of soils. Thus, although these models can be suc-
cessfully applied to studies involving specimen clay minerals, the presence of nat-
urally occurring Fe and Al oxides, and specifically humic substances, severely limits
their application to real world materials.

III. SURFACE CHEMISTRY OF THE SOIL–WATER INTERFACE

The remainder of this chapter will focus on specific examples of water and
solute interactions with soil particles (Plate 2–1). Now that the structural framework
of soil minerals (Chapter 1, Schulze, 2002) and reactive sites of these surfaces (this
chapter) have been presented, several examples illustrating how these sites react
with components of the soil solution will be considered. The intent here is not to
provide a comprehensive review of each process, but to give representative exam-
ples of important constituents of the soil solution and their interaction with soil min-
erals. It should be noted that the interaction of the soil solution with soil minerals
will generally involve more than a single type of active site, and that the division
and assignment of sites is somewhat artificial. Nonetheless, it is hoped that the con-
ceptual framework of active sites will promote increased understanding of the struc-
ture and reactivity of soil minerals.

A. Soil Solution

The soil solution is defined as the aqueous liquid phase of soil at field mois-
ture contents (Wolt, 1994). Chemical reactivity in soils is dominated by reactions
that occur at the soil solution–soil particle interface. Conversely, the chemical
composition of the soil solution reflects the dynamic chemical and physical
processes that occur in the soil. One unique aspect of the surface chemistry of soils
compared with streams and sediments is that the mass of the soil is large relative
to the small amount of water associated with this surface. This can be represented
by the solid phase/water ratio and is represented by the symbol W. A typical W, for
SURFACE CHEMISTRY OF SOIL MINERALS 59

a soil with 15% moisture content, is 4 to 6 kg dm-3. In comparison, W values for


solids suspended in streams and oceans have values ranging as low as a few mi-
crograms per decimeter, nine orders of magnitude lower! Simply stated, because
of their small size, soil particles are characterized by a large, reactive surface in con-
tact with a small volume of soil solution (see question “How thick is the water film
surrounding soil particles?”). In practical terms, the soil solution is strongly impacted
by the high specific surface area of soil particles it contacts. Many laboratory sorp-
tion studies of metals and organics on soil constituents are conducted with W val-
ues many orders of magnitude less than field conditions. Consequently, experimental
data obtained with small W values may not accurately reflect the surface behavior
of soil particles.
Chemical weathering of soil particles is accomplished by the soil solution
(Plate 2–1) that contains reactive species that are brought into contact with the soil
surface. Additionally, products and reactants of biochemical processes occurring
in soil are in close contact with the soil solution. In this context, soils can be con-
sidered to function as massive chromatography columns or chemical reactors
(Richter, 1987). Furthermore, solution phase reactivity is moderated by the processes
shown in Plate 2–1, including partitioning of nonpolar lipophilic (i.e., having a strong
attraction for fats) molecules into biomass or soil organic matter, volatilization to
soil atmosphere, or retention in the soil solid phase by either exchange or precipi-
tation. The main point here is that analysis of the soil solution, although somewhat
difficult to obtain, provides an important compositional “snapshot” that reflects the
dominant chemical, hydrological, biological, and geological processes occurring
in the soil.

B. Water at the Soil–Solution Interface

It has been known for many years that the behavior of water near clay sur-
faces is different from that of bulk water (Mooney et al., 1952a, b). Researchers have
found that water near the surface of expandable clay minerals, for example, is or-
ganized largely by the exchangeable cations and has molecular properties that are
distinct from water molecules in aqueous solution (Sposito & Prost, 1982; John-
ston et al., 1992; Berend et al., 1996). One practical result is that the amount of water
sorbed by expandable clay minerals is controlled by the nature of the exchangeable
cations. Water sorption isotherms of a Wyoming bentonite exchanged with Ca2+,
Ba2+, Na+, and Cs+ are plotted as a function of the relative vapor pressure of water
(analogous to relative humidity) in Fig. 2–4. First, the amount of water sorbed by
these clays is strongly influenced by the exchangeable cation. The amount of water
sorbed follows the hydration energies (Table 2–2) of these cations: Ca2+ > Ba2+ >
Na+ > Cs+. In addition, the plot shows that all of the clays have the potential to sorb
between 20 and 40% of their dry mass in water at high relative humidity, which is
a remarkable and important feature of clay chemistry. As illustrated in Fig. 2–2d,
the hydration characteristics of the exchangeable cations play a major role in or-
ganizing water molecules near the surface of permanently charged clays.
Water molecules surrounding exchangeable metal cations may be strongly po-
larized by their close proximity to the exchangeable metal ions. As the charge of
the cation increases and its size decreases, surface Brønsted acidity increases (first
60 JOHNSTON & TOMBÁCZ

Fig. 2–4. Water sorption isotherms for a Wyoming montmorillonite clay saturated with Ca2+, Ba2+, Na+,
and Cs+ (data from Mooney et al., 1952ab).

demonstrated by Fripiat et al., 1962; Farmer & Mortland, 1966; Mortland, 1967).
Brønsted acids are defined as chemical species capable of donating protons in chem-
ical reactions. For cations with intermediate to large ionic potentials (defined as Z2/r,
where Z is the valence and r is the ionic radius of the ion), the water molecules co-
ordinated to the cation can function as a source of Brønsted acidity. These polar-
ized water molecules surrounding exchangeable cations and undercoordinated sur-
face atoms can donate protons to adjacent solutes more readily than bulk water and
can promote a variety of chemical reactions (Voudrias & Reinhard, 1986). Brøn-
sted acidity of clay mineral surfaces can be quite high (Mortland, 1967; Helsen,
1982), with surface acidities approaching that of concentrated sulphuric acid
(Helsen,1982). Surface Brønsted acidity on montmorillonite follows the order: Al3+
> Mg2+ > Ca2+ > Na+. The implications of these findings are that the surfaces of
clay minerals become highly acidic at low water content. In other words, these types
of surfaces have an increased ability to donate protons to surface species. For ad-
sorbed organic molecules, these conditions often result in the abiotic degradation
of the organic species (Mortland, 1970).

C. Electrified Interface—Diffuse Double Layer Theory

What is the distribution of ions near the charged surfaces of soil minerals?
Knowing the answer to this question allows one to better understand a number of
important processes occurring in soils, including retention of plant nutrients, move-
ment of water and entrained solutes through soil, development of soil structure, and
erosion. For nearly a century, soil researchers have been interested in measuring
and modeling the distribution of charged species near soil particles. Unfortunately,
the distribution cannot be measured directly, and mathematical models are the only
SURFACE CHEMISTRY OF SOIL MINERALS 61

practical way to gain insights into the distribution ionic species near soil minerals.
Among the oldest and yet most widely used models used in soil chemistry to pre-
dict the distribution of ionic species near a charged surface, such as an expandable
clay surface, is the Gouy–Chapman or diffuse double layer theory (van Olphen,
1977). This model was applied initially to describe the charge density of constant
charge surfaces, such as montmorillonite, and was later extended to conditionally
charged surfaces. In this model, the surface is assumed to a be a uniform plane with
a fixed surface potential. For montmorillonite, the surface potential is directly re-
lated to the number of charged siloxane ditrigonal cavities per unit area of surface
(Singh & Uehara, 1986). Depending on the layer charge of the clay mineral, val-
ues typically range from 1 to 0.25 sites nm-2. The ions are treated as point charges
and interact only through electrostatic forces. In addition, the aqueous phase in con-
tact with the surface is (falsely) assumed to have a fixed dielectric constant. Using
these approximations, Gouy (1910) and Chapman (1913) developed a model that
predicts the local concentrations of ions as a function of distance from the surface.
A complete derivation of the Gouy–Chapman model is found in Singh and Uehara
(1986) and will not be developed here beyond the basic working equations. The
Gouy–Chapman model allows one to calculate the local concentration of ions in
solution as a function of distance from a charged surface using the equation:

-zi Fyx
cix = ci¥ exp ______ [4]
‰ RT 
where cix is the concentration of the ion i at a distance x from the surface in moles
per cubic decimeter, ci¥ is the concentration of ion i at an infinite distance from the
surface (i.e., the bulk concentration) in moles per cubic decimeter, z is the valence
of the ion i, F is the Faraday constant (9.648 × 104 C mol-1), R is the molar gas con-
stant (8.314 J mol-1 K-1), and T is the temperature in degrees Kelvin. The yx term
in the exponent is the electrical potential at a distance of x from the surface.
According to Gouy–Chapman theory, the potential at the surface yo (mea-
sured in volts) is related to the surface charge density, sp (measured in coulombs
per square meter) by the following expression (Eq. [5]).

zF y
sp = (8RT eeo c × 103)1/2 sinh _____o [5]
‰ 2RT 
where e is the dielectric constant of water (78.5 at 298 K), eo is the permittivity of
free space (8.854 × 10-12 C2 J-1 m-1), c is the molar electrolyte concentration of
the bulk solution (mol dm-3), and z is the valence of the ion. This equation is valid
for a symmetrical electrolyte where the number of anions and cations in the salt are
equal to each other.
To illustrate the application of Gouy–Chapman theory, consider the mont-
morillonite clay in 0.001 M NaCl presented above. The density of exchange sites
on a representative montmorillonite was 1 × 10-6 sites m-2, which corresponds to
a surface charge density of -0.096 C m-2. Using this value, Eq. [2] becomes

sp = -0.096(C/m2) = 0.00371[sinh(19.4yo)] [6]


62 JOHNSTON & TOMBÁCZ

This equation can then be solved for the potential at the surface. For these condi-
tions, yo has a value of 0.203 V. The potential at a given distance from the surface
is represented by Eq [7].

y(x) = yo exp(-kx) [7]

where yo is the potential at the surface, x is the distance from the surface in me-
ters, k is the Debye Huckel term (m-1) defined in Eq. [8]:

2F 2I × 103 1/2
k= ________ [8]
‰ eeo RT 
where I is the ionic strength in moles per decimeter. Combining the above equa-
tions, Eq. [9] can be used to determine the concentration of anions and cations as
a function of distance from the surface.

-zi Fyo exp (-kx)


cix = ci¥ exp ______________ [9]
— RT 
For example, consider an aqueous suspension of Na-montmorillonite in a so-
lution of 0.001 M NaCl (I = 0.001 M) at 298 K. The value of k for a 0.001 M NaCl
solution is 1.03 × 108 m-1. The double layer thickness is given by the term 1/k and
has a value of 9.6 nm for this solution. In other words, the double layer thickness
for this system extends out about 30 layers of water molecules beyond the surface.
Using this k value, the concentration dependence can be calculated as shown
below.

(-1)(96500)(-0.203) exp (-1.03 × 108 x)


cNa (x) = (0.001) exp ________________________________ [10]
(8.314)(298)

The important features of Gouy–Chapman theory are that it provides a model


to describe the surface potential and local concentrations of ions near a charged sur-
face. This is illustrated for the Na-montmorillonite clay suspension in 0.001 M NaCl
shown in Fig. 2–5. The lower curve represents the concentration of Cl- ions as a
function of distance from the clay surface. Because the montmorillonite surface has
a negative charge, the anions are repelled from the surface. Upon approaching the
surface, the concentration of Cl- sharply decreases within the diffuse double layer
region shown by the dashed vertical line. In contrast, the upper curve represents the
concentration dependence of the Na+ cations as a function of distance from the sur-
face. In addition to constant charge surfaces, such as montmorillonite, Gouy–Chap-
man theory has been successfully applied to variable charge surfaces.
The Gouy–Chapman model is useful to qualitatively predict the concentra-
tions of ions near surfaces. One of the main limitations of the model is that ions are
treated as point charges. As a result, there is no limit to the number of counter ions
that can accumulate near a charged surface and at distances very close to the sur-
face, and the model greatly overestimates the local concentration of counterions.
In addition, because the model treats ions as point charges, it is incapable of pre-
SURFACE CHEMISTRY OF SOIL MINERALS 63

Fig. 2–5. Concentration of Na+ (upper curve) and Cl- (lower curve) plotted as a function of the distance
from a montmorillonite particle suspended in a solution of 0.001 M NaCl with a charge density on
the montmorillonite of 85 cmolC kg-1 and a surface area of 800 m2 g-1. The thickness of the electri-
cal double layer is shown by the dashed line and corresponds to 1/k, which is 9.6 nm for these con-
ditions.

dicting differences between the behavior of ions of the same valence, such as Na+
and Cs+, or Ca2+ and Mg2+.
At the present time, more sophisticated macroscopic and molecular models
are being applied to better understand the surface chemistry of soil minerals. For
a good introduction to current research in chemical speciation modeling, see Davis
and Kent (1990), and for application of these models to soil chemistry, see Loep-
pert et al. (1995). These models are based on a more accurate understanding of the
reactive features on soil minerals and their behavior in soil solution. However, it
is important to remember that these models are based on assumptions about the sur-
face chemistry of the soil surface. As the models become more complex, the num-
ber of fitting parameters has also increased, and thus, one must use these models
with caution.
One of the main prerequisites for the successful application of these models
is to have accurate structural data on the soil minerals. Some of the models them-
selves can be used to correctly predict the structures of soil minerals. Using ab ini-
tio quantum chemical methods, Cygan and coworkers successfully modeled the
structure of kaolinite, for example (Hobbs et al., 1997). Their model accurately rep-
resented the kaolinite structure when compared with the best-available experimental
data. For larger problems, a lower level of theory can be used. For example, Chang
and coworkers (Chang et al., 1997) obtained good agreement with such models and
various structural and thermodynamic properties of expandable clay minerals in
aqueous electrolyte solutions. These models are based on Monte Carlo and mole-
cular dynamic simulations and are being applied to a broad spectrum of problems.
Using similar models, Teppen and coworkers (Teppen et al., 1997, 1998) studied
the interaction of organic compounds with clay surfaces. One of the products of these
64 JOHNSTON & TOMBÁCZ

efforts are molecular models that provide researchers new tools to develop a pre-
dictive understanding about the structure and reactivity of soil minerals.

D. Types of Surface Complexes

Depending on the size, charge, and chemical properties of solutes contained


in the soil solution, several different types of surface complexes can form at the soil
solution–soil particle interface. Surface complexes are formed when a solute (ion
or neutral molecule) contained in the soil solution (Plate 2–1) reacts with the soil
surface to form a stable complex. The three common types of surface complexes
occurring in soils, inner-sphere, outer-sphere, and diffuse layer complexes, are de-
scribed below.

1. Outer-Sphere Complexes and Diffuse Layer Complexes


Outer-sphere and diffuse layer surface complexes are formed between reac-
tive surface sites and a hydrated ion. A shell of water molecules completely sur-
rounds the ion and limits the approach of the hydrated ion to the particle surface to
distances >0.3 nm (about the diameter of a water molecule). A representative
outer-sphere complex of a hydrated cation near a clay surface is shown on the left
side of Plate 2–9. Diffuse layer complexes are similar to outer-sphere complexes
but are situated further from the particle surface (>0.6 nm) as shown on the right
side of Plate 2–9. Using the diameter of a water molecule (~0.3 nm) as a “molec-
ular yardstick”, the thickness of the diffuse layer extends from about 2 to 30 lay-
ers of water molecules from the particle surface. Examples of outer-sphere and dif-
fuse layer complexes include hydrated Ca2+, Mg2+, and Na+ ions near a clay surface,
interlayer Mg2+ ions in vermiculite, and hydrated Cl-, NO3-, ClO4-, SO42-, and
SeO42- ions near positively charged surfaces.

2. Inner-Sphere Complexes
Formation of inner sphere complexes occurs when ions bind directly to the
surface with no intervening water molecules. These types of surface complexes are
restricted to ions that have a high affinity for surface sites and include specifically
adsorbed ions that can bind to the surface through covalent or Coulombic (i.e., elec-
trostatic) interactions. A representative inner-sphere complex is shown in Plate 2–10,
where a phosphate ion is bound to the surface of goethite, FeOOH, through two co-
valent bonds. The most reactive groups on goethite are the singly coordinated Fe–O
groups (referred to as Type A Fe–OH groups), and these groups form a direct, co-
valent bond with the PO43- species. The O atoms of these surface-exposed Fe–O
bonds carry a partial charge and are representative of inorganic surface hydroxyl
groups as discussed in Section II.A.3 above. This complex is called an inner-
sphere complex because there are no water molecules between the adsorbent (i.e.,
goethite) and the adsorbate (i.e., phosphate). More recently, similar reaction mech-
anisms have been proposed for arsenate and selenate complexes with Fe and Al ox-
ides.
Another type of inner sphere complex that occurs in soil minerals is interlayer
K+ ions in mica. In this case, the K+ ions are resident in the interlayer region of the
SURFACE CHEMISTRY OF SOIL MINERALS 65

mica particle to maintain electroneutrality of the clay. The nonhydrated ionic ra-
dius of the K+ ion is close to the dimension of the siloxane ditrigonal cavity (Plate
2–6). Examples of inner-sphere complexes include specific adsorption of oxyan-
ions, such as PO43-, SeO4, and AsO4 with inorganic surface hydroxyl groups, and
interlayer K+ ions in muscovite.
Inner-sphere surface complexes form stronger bonds with soil particles com-
pared with outer-sphere and diffuse layer complexes. As a result, surface reactions
involving the formation of inner-sphere complexes are often irreversible, and the
kinetics of inner-sphere surface complex formation tend to be slower. One impor-
tant example is the fixation of P in highly weathered tropical soils. The central part
of Brazil, for example, is an area known as the Cerrado and covers an area of 125
million hectares. The dominant soils found in the region are acid-Savannah Oxisols,
which are dominated by the presence of Fe and Al oxides. Phosphate has a very high
affinity for Al–OH and Fe–OH groups and can be covalently bound to Al–OH and
Fe–OH groups. When this occurs, the P in the soil, even if present in appreciable
concentration, is effectively fixed in a form that is not available to the plant.

IV. SUMMARY

This chapter has focused on developing an improved understanding of the re-


active nature of soil surfaces on the basis of a knowledge of the surface chemistry
of soil minerals and soil organic matter. We view this topic as a direct extension of
the structural concepts presented in Chapter 1 (Schulze, 2002). The reactivity of
soil minerals is intrinsically related to their heterogeneity, small size, and charged
nature. At the same time, however, these features also present formidable challenges
to soil chemists faced with studying soil minerals and chemical processes occur-
ring in soil and subsurface environments. As information about the surface chem-
istry is gained by using increasingly more powerful methods and as technological
advances improve our ability to analyze difficult samples, experimentalists will be
provided with increasingly powerful tools to examine the surface chemistry of nat-
urally occurring soil particles.

REFERENCES
Adamson, A.W. 1989. Adsorption of gases and vapors in solids. p. 565–570. In Physical chemistry of
surfaces. John Wiley, New York, NY.
Aiken, G.R., D.M. McKnight, R.L. Wershaw, and P. MacCarthy. 1985. Humic substances in soil, sed-
iment, and water. John Wiley & Sons, New York, NY.
Benjamin, M.M., and J.O. Leckie. 1981. Conceptual model for metal–ligand–surface interactions dur-
ing adsorption. Environ. Sci. Technol. 15:1050–1057.
Berend, I., J.-M. Cases, M. Francois, J.P. Uriot, L. Michot, A. Masion, and F. Thomas. 1996. Mecha-
nism of adsorption and desorption of water vapor by homoionic montmorillonites: 2. The Li+
Na+, K+, Rb+ and Cs+-exchanged forms. Clays Clay Miner. 43:324–336.
Bolt, G.H., and M.G.M. Bruggenwert. 1978. Soil chemistry. A. Basic elements. Developments in Soil
Science. 5A. Elsevier, Amsterdam, the Netherlands.
Buffle, J. 1988. Complexation reactions in aquatic systems: An analytical approach. Ellis, Chichester,
UK.
Chang, F.R.C., N.T. Skipper, and G. Sposito. 1997. Monte Carlo and molecular dynamics simulations
of interfacial structure in lithium-montmorillonite hydrates. Langmuir 13:2074–2082.
66 JOHNSTON & TOMBÁCZ

Chapman, D.L. 1913. A contribution to the theory of electrocapillarity. Philos. Mag. 25(6):475–481.
Davis, J.A., and D.B. Kent. 1990. Surface complexation modeling in aqueous geochemistry. p. 177–260.
In M.F. Hochella and A.F. White (ed.) Mineral–water interface geochemistry. Reviews in Min-
eralogy. Vol. 23. Mineralogical Soc. of America, Washington, DC.
Farmer, V.C., and M.M. Mortland. 1966. Infrared study of the coordination of pyridine and water to ex-
changeable cations in montmorillonite and saponite. J. Chem. Soc. 1996A:344–351.
Fripiat, J.J., A. Servias, and A. Leonard. 1962. Adsorption des amines par montmorillonites. Bull. Soc.
Chim. Fr. p. 635–644.
Gouy, G. 1910. Sur la constitution de la charge eletrique a la surface d’un electrolyte. Ann. Phys. (Paris)
9(4):457–468.
Haderlein, S.B., and R.P. Schwarzenbach. 1993. Adsorption of substituted nitrobenzenes and nitrophenols
to mineral surfaces. Environ. Sci.Technol. 27:316–326.
Helsen, J. 1982. Clay minerals as solid acids and their catalytic properties. J. Chem. Educ. 59:1063–1065.
Hobbs, J.D., R.T. Cygan, K.L. Nagy, P.A. Schultz, and M.P. Sears. 1997. All-atom ab initio energy min-
imization of the kaolinite crystal structure. Am. Mineral. 82:657–662.
Huang, P.M. 1990. Organo-alumino polymer associations and their significance in soil and environmental
sciences. p. 179–237. In M.F. De Boodt et al. (ed.) Soil colloids and their associations in ag-
gregates. Plenum, New York, NY.
Huheey, J.E. 1978. Inorganic chemistry. Principles of structure and reactivity. 2nd ed. Harper and Row
Publ., New York, NY.
Jenne, H. 1980. The soil resource: Origin and behavior. Springer Verlag, New York, NY.
Jensen, W.B. 1978. The Lewis acid-base definitions: A status report. Chem. Rev. 78:1–22.
Johnston, C.T. 1996. Sorption of organic compound on clay minerals: A surface functional group ap-
proach. In B. Sawhney (ed.) Organic pollutants in soil environments. Clay Minerals Society, Boul-
der, CO.
Johnston, C.T., and G. Sposito. 1987. Disorder and early sorrow: Progress in the chemical speciation
of soil surfaces. p. 89–100. In L.L. Boersma (ed.) Future developments in soil science research.
SSSA, Madison, WI.
Johnston, C.T., G. Sposito, and W.L. Earl. 1993. Surface spectroscopy of environmental particles by
Fourier transform infrared and nuclear magnetic resonance spectroscopy. p. 1–36. In J. Buffle
and H.P. Van Leeuwen (ed.) Environmental particles. Vol. 2. Environmental Analytical and Phys-
ical Chemistry Series. Lewis Publ., Boca Raton, FL.
Johnston, C.T., G. Sposito, and C. Erickson. 1992. Vibrational probe studies of water interactions with
montmorillonite. Clays Clay Miner. 40:722–730.
Koopal, L.K. 1996. Ion adsorption on mineral oxide surfaces. p. 757–796. In A. Dabrowski and V.A.
Tertykh (ed.) Adsorption on new and modified inorganic sorbents studies in surface science and
catalysis. Vol. 99. Elsevier, Amsterdam, the Netherlands.
Koopal, L.K., and W.H. van Riemsdijk. 1989. Electrosorption on random and patchwise heterogeneous
surface: Electrical double layer effects. J. Colloid Interface Sci. 128:188–200.
Laird, D.A., E. Barriuso, R.H. Dowdy, and W.C. Koskinen. 1992. Adsorption of atrazine on smectites.
Soil Sci. Soc. Am. J. 56:62–67.
Laird, D.A., and B.L. Sawhney. 2002. Reactions of pesticides with soil minerals. p. 765–794. In J.B.
Dixon and D.G. Schulze (ed.) Soil mineralogy with environmental applications. SSSA Book Ser.
7. SSSA, Madison, WI.
Loeppert, R.H., A.P. Schwab, and S. Goldberg. 1995. Chemical equilibrium and reaction models.
SSSA, Madison, WI.
Mooney, R.W., A.G. Keenan, and L.A. Wood. 1952a. Adsorption of water vapor by montmorillonite.
II. Effect of exchangeable ions and lattice swelling as measured by x-ray diffraction. J. Am. Chem.
Soc. 74:1371–1374.
Mooney, R.W., A.G. Keenan, and L.A. Wood. 1952b. Adsorption of water vapor by montmorillonite. I.
Heat of desorption and application of BET theory. J. Am. Chem. Soc. 74:1367–1374.
Mortland, M.M. 1967. Protonation of compounds at clay mineral surfaces. p. 691–699 In J.W. Holmes
(ed.) International Congress of Soil Science Transactions, 9th. Adelaide, Australia. Vol. 1.
ISSS. American Elsevier, New York, NY.
Mortland, M.M. 1970. Clay–organic complexes and interactions. Adv. Agron. 22:75–117.
Newman, A.C.D., and M.H.B. Hayes. 1990. Some properties of clays and of other soil colloids and their
influences on soils. p. 39–55. In M.F. De Boodt et al. (ed.) Soil colloids and their associations
in aggregates. Plenum, New York, NY.
Oades, J.M., 1989. An introduction to organic matter in mineral soils. p. 89–159 In J.B. Dixon and S.B.
Weed (ed.) Minerals in soil environments. SSSA Book Ser. 1. SSSA, Madison, WI.
SURFACE CHEMISTRY OF SOIL MINERALS 67

Richter, J. 1987. The soil as a reactor. Modeling processes in the soil.Catena Verlag, Cremlingen, Ger-
many.
Sanchez, R.M.T., P.G. Rouxhet, W.E.E. Stone, and A.J. Herbillon. 1990. Interaction of stable and
metastable monomeric iron (III) species with a kaolinitic soil clay. p. 105–117. In M.F. De Boodt
et al. (ed.) Soil colloids and their associations in aggregates. Plenum, New York, NY.
Shannon, R.D. 1976. Revised effective ionic radii and systematic studies of interatomic distances in
halides and chalcogenides. Acta Cryst. A32:751–787.
Schulten, H.R., and M. Schnitzer. 1995. Three-dimensional models for humic acids and soil organic mat-
ter. Naturwissenschaften 82:487–498.
Schulten, H.R., and M. Schnitzer. 1997. Chemical model structures for soil organic matter and soils.
Soil Sci. 162:115–130.
Schulze, D.G. 2002. An introduction to soil mineralogy. p. 1–36. In J.B. Dixon and D.G. Schulze (ed.)
Soil mineralogy with environmental applications. SSSA Book Ser. 7. SSSA, Madison, WI.
Schwertmann, U. 1990. Some properties of soil and synthetic iron oxides. p. 57. In M.F. De Boodt et
al. (ed.) Soil colloids and their associations in aggregates. Vol. 215. Plenum, New York, NY.
Singh, U., and G. Uehara. 1986. Electrochemistry of the double layer: Principles and applications to
soils. p. 1–38. In D.L. Sparks (ed.) Soil physical chemistry. CRC Press, Boca Raton, FL.
Sposito, G. 1984. The surface chemistry of soils. Oxford University Press, New York, NY.
Sposito, G., and R. Prost. 1982. Structure of water adsorbed on smectites. Chem. Rev. 82:553–573.
Stevenson, F.J. 1994. Humus chemistry: Genesis, composition, reactions. 2nd ed. Wiley, New York, NY.
Teppen, B.J., K. Rasmussen, P.M. Bertsch, D.M. Miller, and L. Schafer. 1997. Molecular dynamics mod-
eling of clay minerals. I. Gibbsite, kaolinite, pyrophyllite, and beidellite. J. Phys. Chem.
101:1579–1587.
Teppen, B.J., C.H. Yu, D.M. Miller, and L. Schafer. 1998. Molecular dynamics simulation of sorption
of organic compounds at the clay mineral aqueous solution interface. J. Comput. Chem.
19:144–153.
Tombacz, E.M. 1990. The possibility of heterocoagulation between montmorillonite and humic sub-
stances. Appl. Clay Sci. 5:265–270.
Van Olphen, H. 1977. An introduction to clay colloid chemistry. 2nd ed. Wiley Interscience, New York,
NY.
Voudrias, E.A., and M. Reinhard. 1986. Abiotic organic reactions at mineral surfaces. American Chem-
ical Society, Washington, DC.
Wershaw, R.L. 1993. Model for humus. Environ. Sci. Technol. 27:814–816.
Wolt, J.D. 1994. Soil solution chemistry: Applications to environmental science and agriculture. John
Wiley, New York, NY.
Yamagishi, A. 1993. Chirality recognition by a clay surface modified with an optically active metal
chelate. p. 307–347. In K. Tamaru (ed.) Dynamic processes on solid surfaces. Plenum Press, New
York, NY.
Yariv, S. 1992. Wettability of clay minerals. p. 279–326. In M.E. Schrader and G. Loeb (ed.) Modern
approaches to wettability: Theory and applications. Plenum Press, New York, NY.
Yatsimirksii, K.B., and V.P. Vasil’ev. 1960. Instability constants of complex compounds. Pergamon, Elms-
ford, NY.
Zachara, J.M., S.C. Smith, J.P. McKinley, and C.T. Resch. 1993. Cadmium sorption on specimen and
soil smectites in sodium and calcium electrolytes. Soil Sci. Soc. Am. J. 57:1491–1501.

Vous aimerez peut-être aussi