You are on page 1of 245

The Surface Chemistry of Soils

Garrison Sposito



THE SURFACE CHEMISTRY


OF SOILS

THE SURFACE CHEMISTRY


OF SOILS

Garrison Sposito
University of California, Riverside

Oxford University Press New York


Clarendon Press Oxford
1984

Copyright

1984 by Oxford University Press, Inc.

Library of Congress Cataloging in Publication Data


Sposito, Garrison, 1939The surface chemistry of soils.
Includes index.
1. Soil chemistry. I. Title.
S592.5.S658 1984
631.4'1
ISBN 0-19-503421-X

Printing (last digit):

84-3936

98 7 65 4 3

Printed in the United Stllte!! of Americn

To Our Parents
Albert and Geraldine Sposito
and
Ernest and Virginia Campbell

PREFACE

The surface chemistry of naturally occurring solids was initiated in its


modern sense by soil chemists about one hundred five years ago. Jakob
Maarten van Bemmelen, in his studies on soils sampled from polders in
The Netherlands, published the first solute adsorption isotherm, proposed
the first adsorption isotherm equation (now referred to as the Freundlich
isotherm), and distilled from his results that the adsorptive powers of
ordinary soils depend on the colloidal silicates, humus, silica, and iron
oxides they contain. This profound first effort established one of the two
great leitmotivs in the surface chemistry of soils: the use of adsorption
isotherm equations to characterize equilibria between constituents of the
soil solution and the solid phases in soil. The second great theme was
developed thirty-five years later in the papers of Robert Gans (Ganssen),
who suggested that ion exchange reactions in soils could be described by a
mass-action coefficient (conditional equilibrium constant). Once the concept of thermodynamic activity was developed, the rigorous extension of
the ion exchange mass-action coefficient to an exchange equilibrium
constant became possible. This last foundational step was taken by Albert
P. Vanselow, a student of G. N. Lewis.
During the past fifty years, surface chemical phenomena in soils have
been interpreted within the classical conceptual framework created by van
Bemmelen, Gans, and Vanselow. The successes have been many, but with
the growth of applications has come a deepening of understanding,
particularly in respect to the limitations of the classical approach. It is
apparent now that neither the adsorption isotherm nor the exchange
equilibrium constant provides any unique information as to the mechanisms of surface chemical reactions. Adsorption isotherm equations bear
no simple relationship to the chemical equilibria they model, since it is
possible to derive the same equation under different sets of mutually
exclusive hypotheses about the adsorption mechanism. The isotherm
equation is best regarded as II curve-fitting expression with parameters

viii

PREFACE

having potential correlative and predictive value. In a similar vein, the


thermodynamic description of ion exchange is known now to be a macroscopic theory of mixtures that is independent of exchange mechanisms.
Surface chemistry and ion exchange have no necessary connection, as the
many documented geochemical coprecipitation reactions-which also are
ion exchange reactions-illustrate so clearly.
The mechanisms of surface chemical reactions represent a problem in
coordination chemistry, which is the study of complexes, molecular units
comprising a central group surrounded by other atoms in close association.
This book is principally an introduction to the interpretation of surface
phenomena in soils from the point of view of coordination chemistry.
Therefore the basic concept to be discussed is the surface functional group,
the central moiety in surface complexes, whose formation provides the
most important mechanism of adsorption by the solid phases in soils. No
detailed consideration of adsorption isotherm equations or the thermodynamic theory of ion exchange is presented, except insofar as their
tenuous relation with surface coordination chemistry is to be illustrated.
The discussion in this book is intended to be self-contained, but a previous
exposure to soil physical chemistry, soil mineralogy, and the fundamentals
of inorganic chemistry will prove helpful.
The first chapter of this book reviews three basic experimental divisions
of the surface chemistry of soils: the nature of adsorbing solid phases, the
measurement of specific surface area, and the determination of surface
charge density. Emphasis is placed on the view that the adsorbing solids in
soils are inorganic and organic polymers bearing surface functional groups
whose reactivity determines the operational meanings of surface area and
surface charge. The second and third chapters provide a general description of the solid-aqueous solution interface in soils which does not rely on
any detailed molecular model of the interfacial region, such as diffuse
double layer theory. The intent in these chapters is to introduce familiar
interfacial phenomena as much as possible in a model-independent context
while preserving the guiding principles of coordination chemistry. The
fourth and fifth chapters deal with the experimental and theoretical aspects
of adsorption by soil constituents. In these chapters, mechanisms of adsorption as examples of surface complexation are discussed and molecular
models of the interfacial region are described. The complementary roles
of the diffuse ion swarm and of surface complexes are examined through a
survey of recent chemical models of adsorption. The sixth chapter presents
a brief review of the relationship between surface coordination chemistry
and soil colloidal phenomena. This chapter can serve as an introduction to
the key concepts of colloid structure and interparticle forces in soil clay
suspensions.
The initial inspiration to develop this monograph has come from the very
beautiful research of Dr. Werner Stumm and his colleagues on the surface
chemistry of hydrous oxides. The development of the coordination chemistry approach to adsorption. pioneered hy him and Dr. Paul Schindler some

PREFACE

ix

fifteen years ago, has made a truly seminal contribution to the study of
natural water systems. Those of us fortunate enough to have known him
and his work share in the reflected light of his fine accomplishments.
I should like to thank also Dr. James J. Morgan for continual encouragement and support in ways too numerous to count. Dr. James O. Leckie has
been most helpful over the past several years as I began to learn the
vicissitudes of computer modeling of surface phenomena. I am grateful
also to Michael Essington for providing me with a set of lecture notes for
my course on soil physical chemistry that has helped me significantly during
the writing of this book. Chapters 1 and 2 were reviewed in draft by Dr.
James P. Quirk, who made many suggestions for their improvement.
Chapter 1 also was read by Dr. Roger L. Parfitt and Chapter 2 by Dr. Rene
Prost, both of whom have helped to clarify my thinking. Chapter 3 was
reviewed by Dr. Robert J. Hunter, who gave me the benefit of his great
experience with electrified interfaces. I must thank also Dr. Clifford T.
Johnston and Dr. Sabine R. Goldberg for their most careful scrutiny of
several chapters in an effort to expunge errors and simplify sentences.
Finally, I express deep gratitude to Diana Deporto and Linda Bobbitt for
their skill in drawing the figures and to Martha Stephans and Sharon
Conditt for their excellent typing of the manuscript.
Riverside, California
August 1983

G.S.

There are ancient cathedrals which, apart from their consecrated purpose,
inspire solemnity and awe, Even the curious visitor speaks of serious things,
with hushed voice, and as each whisper reverberates through the vaulted
nave, the returning echo seems to bear a message of mystery. The labor of
generations of architects and artisans has been forgotten, the scaffolding
erected for their toil has long since been removed, their mistakes have been
erased, or have become hidden by the dust of centuries. Seeing only the
perfection of the completed whole, we are impressed as by some superhuman agency. But sometimes we enter such an edifice that is still partly under
construction; then the sound of hammers, the reek of tobacco, the trivial jests
bandied from workman to workman, enable us to realize that the great
structures are but the result of giving to ordinary human effort a direction
and a purpose,
Gilbert Newton Lewis
and Merle Randall l

G. N, Lewis und M. Rundall, Thermodynamics and till' Free f.'Tll'fNY of Chemical Substances.
Mc(irllwllill Book ('0" New York. 11)2,1 Reprinted with the permission or the publisher.
I

CONTENTS

1. THE REACTIVE SOLID SURFACES IN SOILS

1.1. Structural Chemistry of the Surface-Reactive Solid Phases


Soils, 1
1.2. Surface Functional Groups in Soil Clays, 12
1.3. Solid Surfaces in Natural Soils, 19
1.4. Specific Surface Area, 23
1.5. Surface Charge Density, 35
2. THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

2.1.
2.2.
2.3.
2.4.

Liquid water, 47
Electrolyte Solutions, 54
Water near Phyllosilicate Surfaces, 57
The Solvent Properties of Adsorbed Water, 69

3. THE ELECTRIFIED INTERFACE IN SOILS

3.1.
3.2.
3.3.
3.4.
3.5.

The Balance of Surface Charge, 78


Points of Zero Charge, 81
Potentials near an Electrified Interface, 88
Electrokinetic Phenomena, 94
Negative Adsorption, 106

4. INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

4.1.
4.2.
4.3.
4.4.
4.5.

The Adsorption Isotherm, 113


Adsorption Versus Precipitation, 122
Metal Cation Adsorption, 128
Inorganic Oxyanion Adsorption, 138
Organic Matter Adsorption, 143

ill

xii

CONTENTS

5. CHEMICAL MODELS OF SURFACE COMPLEXATION

5.1. The Diffuse Double Layer Model, 154


5.2. Surface Complexation Models: Statistical Mechanics, 162
5.3. The Constant Capacitance Model, 169
5.4. The Triple Layer Model, 177
5.5. The Objective Model, 185
5.6. The Structure of Surface Complexation Models, 188
6. SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

6.1. The Smectite Quasicrystal, 198


6.2. Interparticle Forces in Phyllosilicate Suspensions, 205
6.3. The Stability of Soil Colloidal Suspensions, 217
SELECTED PHYSICAL CONSTANTS, 229
SUBJECT INDEX, 231

THE SURFACE CHEMISTRY


OF SOILS

1
THE REACTIVE SOLID SURFACES
IN SOILS

1.1. STRUCTURAL CHEMISTRY OF THE SURFACE-REACTIVE


SOLID PHASES IN SOILS

The solid phases that exhibit surface reactivity in soils are to be found
primarily in the clay fraction. This well-known fact is a consequence of the
geometric relationship between particle volume and surface area: in a
closely packed mass of solid particles, the total surface area increases as the
degree of subdivision of the mass increases. For example, in a cubic meter
of medium sand particles, each assumed spherical with a diameter of
500 JLm, the total surface area is about 1.2 x 104 rrr'. In the same volume
of clay particles, however, each with a diameter of 2 JLm, the total surface
area is about 3 x 106 m 2 , or 250 times that of the sand particles. 1 It is
evident from this comparison that a study of the solid phases in soils in
relation to surface chemistry need focus only on those solids that are
common in clay fractions.
From the point of view of structural chemistry, most of the solid phases
contained in soil clays are polymers, i.e., compounds or mixtures of
compounds formed from the bonding together of repeating molecular
units. If the repeating structure in a solid phase persists throughout a
molecular region whose diameter is at least 3 nm, the solid phase is said to
be crystalline. If structural regularity does not exist over molecular
distances this large, the solid phase is termed amorphous. This distinction
concerning structural regularity is intended principally as a general guide.
Even among the crystalline solids in soil clays there is wide variation in
molecular order, with disorder introduced often by isomorphic substitutions of cations and anions and by the irregular stacking of crystalline
layers.
The most important molecular structural units in the inorganic polymers
found in soil days are the silica tetrahedron, sio1 ,and the octahedral
comprising a mctal calion, M'" I , and six anions, X" .
complex MXg'

r:

TETRAHEDRAL SHEET

DIOCTAHEDRAL SHEET

'J_--X b J+-----M m+

Figure 1.1. Sheet polymeric structure of Si04- and Mx~m~6b). The open circles at
the polyhedral vertices in each perspective view are shown directly below in a
projection along the crystallographic a axis.

THE REACTIVE SOLID SURFACES IN SOILS

Both of these units can polymerize to form sheet structures, as illustrated


in Fig. 1.1. The tetrahedral sheet results from the sharing of corners
occupied by oxygen ions, an arrangement that is optimal for screening the
positive coulomb field produced by the Si4 + ions at the centers of the
tetrahedra. The octahedral sheet is built up through the sharing of edges,
which brings neighboring metal cations nearer to one another than in the
case of shared corners.
If all possible sites for metal cations in an octahedral sheet are occupied,
each cation shares two anions with each of six neighbouring cations and the
sheet is termed trioctahedral. If just two thirds of the possible sites are
occupied, each cation shares two anions with each of three neighboring
cations and the sheet is termed dioctahedral. A dioctahedral sheet is shown
in Fig. 1.1. Since the sharing of edges in an octahedral sheet brings
neighboring metal cations closer together than when octahedral corners
are shared, the cationic coulomb field is not screened as well and the
octahedra tend to become distorted in order to minimize cation repulsion.
The principal feature of this distortion is a shortening of anion-anion
distances along the shared edges in order to provide additional screening of
the cationic electric field?,3
Crystalline solids that contain only silica tetrahedral sheets do not exist
in soils, but carbonate, oxide, oxyhydroxide, and hydroxide solids that have
metal cations in octahedral coordination are widespread. The ratio of the
radius of the common metal cations in soil clays to that of bivalent oxygen
usually ranges between 0.4 and 0.7 (Table 1.1), which means that,
according to the Pauling Rules," octahedral coordination of the metal
cations with 0(- II) is preferred. Because of their great abundance in
the lithosphere and their low solubility in the normal range of soil pH
values, aluminum, iron, and manganese form the most important oxide,

Table 1.1. Ionic radii of elements commonly found in clay fractions of soils

Element

Radius,
nm a

Radius,
ratio"

Li(I)
D( -II)
F(-I)
Na(I)
Mg(lI)
Al(III)
Si(IV)
K(I)
Ca(II)

0.074
0.140
0.133
0.102
0.072
0.053
0.026
0.138
D.IOO

0.53

0.73
0.51
0.38
0.19
0.99
0.71

Element

Radius,
nm a

Radius
ratio"

Co (III)
Mn(III)
Mn(IV)
Fe (II)
Fe (III)
Cu(lI)
Zn(II)
Cd(lI)
Pb(lI)

0.061
0.065
0.054
0.078
0.073
0.073
0.075
0.095
0.118

0.44
0.46
0.39
0.56
0.52
0.52
0.54
0.68
0.84

"Ionic radii for octahedral coordination [excepting Si(IV). which is for tetrahedral coordination] compiled
from R. D. Shunnon und C. D. Prewitt, Effective ionic radii in oxides and fluorides. Acta Cryst. 825: lJ25
( IlJIllJ).
" r( cutinn l/ rl ( )(

Ill!

THE SURFACE CHEMISTRY OF SOilS

oxyhydroxide, and hydroxide minerals in soil clays. These minerals


are listed in Table 1.2, and representative octahedral sheet structures are
shown in Fig. 1.2.
Among the iron compounds listed in Table 1.2, goethite is the one most
often found in soils, regardless of climatic region. Under oxic conditions
and in the presence of iron-complexing ligands that inhibit crystallization,
however, ferrihydrite may precipitate from a soil solution. This poorly
crystalline solid comprises sheets of octahedra with Fe(III) coordinated to
0, OH, and OH z in a defect-sprinkled arrangement similar to that in
hematite. Ferrihydrite can transform either to hematite, which ultimately
transforms to goethite, or directly to goethite. Goethite is the most
thermodynamically stable of the iron oxides and, therefore, is the solid
phase expected finally in soil clays." As indicated in Fig. 1.2, the oxygen
ions in goethite lie in planes perpendicular to the crystallographic a axis
and the iron cations are coordinated in distorted octahedra that share
edges. Some of the octahedral vertex ions are OH- groups, which form
hydrogen bonds with neighboring oxygen ions.
Gibbsite is the most important of the aluminum minerals listed in
Table 1.2. Its molecular structure is illustrated in Fig. 1.2. The dioctahedral sheets in the structure are bound together by hydrogen bonds
between opposed hydroxyl groups. Hydrogen bonding also occurs between
hydroxyl groups along the edges of unfilled octahedra within a sheet,
thereby producing additional distortion of the aluminum octahedra beyond
that caused by the sharing of their edges.f
The most commonly found manganese mineral in soils is birnessite,
with lithiophorite, the other manganese oxyhydroxide listed in Table 1.2,
restricted largely to acidic soils. Birnessite contains sheets of MnO~
octahedra linked in some fashion with Mn(III) , Mn(II), Na(I), and
Ca (II) ions coordinated to both hydroxyl groups and water molecules.
In lithiophorite, sheets of MnO~- octahedra alternate with sheets of
Alo.67Lio.33 octahedra."
The polymeric sheet structures illustrated in Fig. 1.1 combine to form
the phyllosilicate class of minerals. This bonding together of tetrahedral
Table 1.2. Metal oxides, oxyhydroxides, and hydroxides commonly found in clay
fractions of soils

Name
Anatase
Birnessite
Boehmite"
Ferrihydrite
Gibbsite"
Goethite"

Chemical formula

rio,
Nao.7CaO.3Mn70142.8HzO
y-AlOOH
FeZ03' 2FeOOH 2.6H zO
y-Al(OHh
a-FeOOH

Name

Chemical formula

Hematite"
Ilmenite
Lepidocrocite"
Lithiophorite
Maghernite":"
Magnetite"

a- FeZ03
FeTi03
y-FeOOH
(Al,Li)MnOz(OH)z
y-FeZ0 3
FeFez04

The y denotes cubic dme,pllckinll 01' unions, lind the " denotes
h Sllllle Ill' the !-'c(III) Ion. nrc in It'lrllhe<lrul clH'f<llnlllilln,

he~111l1l11l11

dIlNe-pllcking.

0/

Fe 3+

Ii ~cJ-'"

'Xi-p;" d
!~ 9
,1,,"'i

GOETHI TE, a-FeOOH

GIBBSITE, y-AI(OHh

Figure 1.2. The molecular structures of goethite and gibbsite projected along the crystallographic c axis
(upper) and a axis (lower). Hydrogen bonds in goethite are indicated by dashed lines, and an Fe03(OH)3
octahedron is outlined in the a axis projection. (After Brown et at. 7 )

(,

THE SURFACE CHEMISTRY OF SOILS

and octahedral sheets occurs through the apical oxygen ions in the former
and always produces a significant distortion of the anion arrangement in
the final layer structure. The principal distortion is caused by the fact that
the apical oxygen ions in the tetrahedral sheet cannot fit into the vertices
of the octahedra to form a layer and still preserve the ideal hexagonal
pattern of the tetrahedra. In order to fuse the two sheets, pairs of adjacent
tetrahedra must rotate alternately clockwise and counterclockwise by
about 20 around an axis perpendicular to their basal plane. This rotation
lowers the symmetry of the cavities in the basal plane of the tetrahedral
sheet (Fig. 1.1) from hexagonal to trigonal. Besides this distortion, the
sharing of edges in the octahedral sheet and the isomorphic substitution of
the cations in both sheets tend to make the structure thinner and its basal
surfaces less planar (slightly corrugated) than in an unconstrained
polymer. 7
The phyllosilicates in soil clays can be classified into three layer types,
distinguished by the numbers of tetrahedral and octahedral sheets combined, and five groups, differentiated by the kinds of isomorphic cation
substitutions that occur. The layer types are shown in Fig. 1.3, and the
groups are described in Table 1.3. The 1: 1 layer type consists of one
tetrahedral and one octahedral sheet. In soil clays, it is represented by the
kaolinite group, with the unit cell chemical formula [Si4](AI4)OlO(OH)8,
where the cation enclosed in brackets is in tetrahedral coordination and
that enclosed in parentheses is in octahedral coordination. Normally there
is no significant isomorphic substitution for Si(IV) or Al (III) in this group,
and, as is common with soil clay minerals, the octahedral sheet has two
thirds of its cation sites occupied (dioctahedral phyllosilicate).
The 2: 1 layer type has two tetrahedral sheets sandwiching an octahedral
sheet. The three clay groups with this structure are (illitic) mica, vermiculite, and smectite (montmorillonite), each with the general unit cell
chemical formula.f

Table 1.3. Phyllosilicate groups commonly found in clay fractions of soils7 8

Group
Kaolinite
Mica (illitic)
Vermiculite
Smectite"
Chlorite

Layer
type
1:1
2:1
2:1
2: 1

2: ) with
hydroxide
intcrlayer

"nnrl,'lIl1y IIIClllllllllnlllln,I, III ",II rillY'

Layer
charge
<0.01
1.4-2.0
1.2-1.8
0.5-1.2
Variable

Chemical formula
coefficients

6.8

3
3
3.2
8.4

7
8
2.4

c'

c + c' = 0.25

0.5
c + c'

o.s

= 0.2
).5

_--M m +

0 2-

-<:):~~~X~~~X~--Si4+
I: I LAYER

2:1 LAYER

2: I LAYER WITH HYDROXIDE INTERLAYER


Figure 1.3. The three layer types for phyllosilicate structures in soil clays. All
shown here are dioctahedral, with hydroxyl groups shown as shaded circles.

THE SURFACE CHEMISTRY OF SOilS

where M represents 1 mole of cation charge on the basal surfaces; a, b,


and c' are stoichiometric coefficients; and
x = 12 - a - b - c

C,

(1.1)

is the layer charge, the number of moles of net electron charge per unit cell
that is produced by isomorphic substitutions. As indicated in Table 1.3, the
three 2: 1 layer groups differ from one another in two principal ways: the
layer charge decreases in the order mica > vermiculite > smectite and
the vermiculite group is further distinguished from the smectite group
by the extent of isomorphic substitution in the tetrahedral sheet. However,
in the montmorillonite subgroup of the smectites (the one of principal
significance in soil clays), there is a spectrum of isomorphic substitution
patterns ranging from the Chambers type, with (8-a) values up to 0.6, to
the Otay type, with (8-a) values up to 0.18. 7
The 2:1 layer type with hydroxide interlayer is represented by dioctahedral chlorite in soil clays (Figs. 1.3). The unit cell chemical formula of
this mineral can be expressed"
[SiaAIs-a](AlbFe (III)cMgc') 020(0 H) 16
The octahedrally coordinated cations in chlorite reside in two sheets: one
comprising M(OHh04'-lO octahedra (with M'?" = AI3+ , Fe 3 + ,or Mg2+)
sandwiched in a 2: 1 layer and one comprising principally AI(OH)~
octahedra situated on a basal surface of the 2; 1 layer. In order to preserve
the electroneutrality of the whole structure, the octahedral occupancy
often is greater than the expected eight per unit cell for the two dioctahedral sheets; the resulting excess positive charge balances the excess
negative charge created by isomorphic substitution in the tetrahedral
sheets.
Structural disorder in all the soil minerals listed in Tables 1.2 and 1.3 can
be induced through isomorphic substitutions for the principal cations.
The range of these substitutions in the Periodic Table is very broad, as
indicated in Table 1.4. 9 For example, seven metals are known to substitute

Table 1.4. Metals found in soil clay minerals through isomorphic substitution and inclusion

Minerals
Fe oxides
Mn oxides
Ca carbonates
Illites
Vermiculties
Smectites

Substituted metals
AI, V, Mn, Ni, Cu, Zn, Mo

Fe, Co, Ni, Zn, Pb


V, Mn, Fe, Co, Zn, Cd, Pb
Mg, AI, V, Ni, Co, Cr, Zn, Cu, Pb
Mg, AI, Ti, Mn, Fe
Mg, AI, Ti, V, Cr. Mn, Fe, Co, Ni, Cu,
Zn, Ph

THE REACTIVE SOLID SURFACES IN SOILS

for iron in goethite and hematite, with soil goethites containing aluminum
up to concentrations of 32 mole per cent."
A more pronounced structural disorder often exists in freshly precipitated silica or metal hydroxides in soils, since these compounds typically
are amorphous. Structurally disordered aluminosilicates, known collectively as allophane and imogolite, are common in the clay fractions of
soils formed on volcanic ash parent material. 10
The molecular structure of allophane is not well understood, but it is
thought to consist of a 1: 1 phyllosilicate layer riddled with defects (vacant
ion sites) and containing Al in both the tetrahedral and octahedral sheets.
The many defects promote curling of the layer into the form of a hollow
spherule about 5 nm in diameter whose outer boundary contains many
apertures through which small molecules or ions can enter. As this
structural concept implies, allophane often is found in association with
soil clay minerals of the kaolinite group, especially the hydrated polymorph halloysite. Imogolite, which has the general empirical formula
1.1 SiO z"Alz03"2.3-2.8HzO, exhibits a tubular morphology. The tube
unit contains Al only in octahedral coordination and exposes a defective,
gibbsite-like surface.
When one turns to the organic solids in soils, the uncertainties regarding
structural chemistry are even more complex than those encountered with
the amorphous aluminosilicates. It is, in fact, not possible to describe a
developed molecular structure for these compounds at present, although
something can be said about the precursors of such a structure. Collectively, the dark microbially transformed organic solid materials that
persist in soils are termed humic substances, and it is known that their
synthesis involves phenolic compounds resulting from the decomposition of proteins and carbohydrates (Fig. 1.4).11 The phenolic compounds,
whether derived from lignin decomposition or through microbial synthesis,
may polymerize to contribute to biologically stable structures of large
relative molecular mass. Lignin, carbohydrates, and proteins themselves
are organic polymers whose microbial decomposition produces smaller
molecular units that can be bound into humic substances. Lignin is a
polymeric material comprising phenylpropenyl alcohols with one hydroxyl
group (Fig. 1.5). In coniferous lignin, coniferyl alcohol units (shown at
the lower left in Fig. 1.4) bond together in a variety of ways to form a
three-dimensional framework. The carbohydrates are in part polymeric
compounds formed by repeated condensation reactions between hexose
and pentose sugars. In cellulose (Fig. 1.5) for example, the hexose sugar
glucose forms a glucoside (substitution of the hydrogen atom in the
hydroxyl group adjacent to the oxygen atom in the ring) with another
glucose molecule to initiate a polymer structure.
The proteins also are condensation polymers. In this case, the fundamental molecular units are amino acids, which combine according to the
reaction

DEGRADATION OF LIGNIN
(CONIFEROUS LIGNIN)

C-

PHENOLS BY
MICROBIAL SYNTHESIS
ALIPHATIC
CARBON
SOURCE

PROTEINS----......

lY
y

DEGRADAT ION OF
PROTEINS

OCH

f
Rh

PEPTIDES----......
AMINO ACIDS ...............
AMMONIA

I
-CH
I
CH20H

COOH
OH

~Y
OH

RyY0H

R~

oA-(
OH

HO--y

OH . .
RyOH
~~

OH

OH
OH
XOCH 3

CH
II
CH
I
COOH

OTHER PHENOLS
OF PLANTS

OH

DEGRADAT ION OF
CARBOHYDRATES

11

THE REACTIVE SOLID SURFACES IN SOILS

H
H
I II
I II
H-N-C-C
+H-N-C-C
I I 'OH
I I 'OH

R1

1
H I H RI

H-I-N, I /C '" ,I /C, /OH + H 20

C
I

"N

(1.2)

II

where Rand R 1 are arbitrary molecular substituents. The repeating group,


enclosed in a box in Eq, 1.2, is called a peptide. Long polymer chains
comprising these groups often take the form of a right-hand helix with 3.6
peptides per turn and with hydrogen bonds lying approximately parallel to
the helical axis (Fig. 1.6).12 Short pieces of these chains may result from

Figure 1.5. Polymeric structures in coniferous lignin and cellulose.

CONIFEROUS LIGNIN

CELLULOSE

12

THE SURFACE CHEMISTRY OF SOilS

Figure 1.6. Structure of the a-helix peptide chain. The shaded planes illustrate the
orientation of the amide groups, CONH, and interrupted heavy lines denote
hydrogen bonds. (After Walton 12 )

protein degradation in soils. These polypeptides in turn can bond directly


into the structure of humic substances or can react with carbohydrate
degradation products to form amino sugars which are then incorporated
into humic compounds.
It is evident that many chemical and microbiological processes mediate
the polymerization reactions leading to the formation of organic, surfacereactive solids from the precursors shown in Fig. 1.4. This complexity has
defied attempts thus far to develop a unified concept of molecular structure
for humic substances. However, knowledge of the general nature of the
structural precursors makes it possible at least to list the functional groups
expected in the final product. The stereochemistry and reactivity of these
groups when exposed on organic surfaces form the basis of the reactivity of
the surfaces themselves.
1.2. SURFACE FUNCTIONAL CROUPS IN SOIL CLAYS

The surface reactivity of the solid phases in soils derives from the chemical
behavior of surface functional /?,oUP,\' in soil clays. A surface functional
group is a chemically reactive molecular unit hound into the structure of a

THE REACTIVE SOLID SURFACES IN SOILS

13

solid at its periphery such that the reactive components of the unit can be
bathed by a fluid. Surface functional groups may be bound to either
organic or inorganic solids and may exhibit any structural arrangement
conceivable for the functional groups in individual small molecules, but
they differ from ordinary functional groups in two related ways. First,
because they are bound into a solid framework, surface groups cannot be
brought to a state of infinite dilution relative to one another in an aqueous
suspension, as can the groups on small dissolved molecules. Unless the
integrity of the polymer to which they are attached is destroyed, surface
functional groups remain separated by fixed distances regardless of how
dilute the polymer suspension itself may become. Second, the reactivity of
surface functional groups is, in essence, a cooperative property. The fact
that neighbouring groups either have or have not reacted when a solid
surface interacts with the constituents in a fluid phase affects the behavior
of any single group. For example, if an electrically neutral, proton-containing functional group is surrounded by like neighbors that have lost their
protons and become ionized, it is less likely to release its proton than if the
neighboring groups had retained their protons. The reason for this effect is
the negative coulomb field produced by the nearby ionized functional
groups, which exerts a force on the remaining protons to keep them near
the surface.
....
When a surface functional group reacts with a molecule dissolved in a
surrounding fluid to form a stable unit, a surface complex is said to exist
and the formation reaction is termed surface complexation. Two broad
categories of surface complexes can be distinguished on structural groands:
if no molecule of the bathing solvent is interposed between the surface
functional group and the molecular unit it binds, the complex formed is
called an inne~:~P!lq!!_c:.g,!!!ulex; if at least one solvent molecule is inter\ posed between the functional group and the bound molecule, the complex
formed is called an q!:!!ef.:J2h:f!,r.(C;Q1Jlplc~. As a general rule,. outer-sphere
surface complexes jnyolve elect~()staticb2,nal~g!!i~~hal1i~m~'~n!:lJhe.r.e.!ore
, a.re ..less . stable .!~':lp._i~~~-sI?h.~rt:_~.!!rt~tL9.lIlJ2Ie.~es, w~!~1:1 . . J.lec~ssai-il y
;. involve either ionic or covalent QQnciing 9Ls()It1ecQITl9i~.~!i~I! l:)ftheiwo.
It is evident from these definitional remarks that the surface chemistry of
a soil is determined in large measure by the nature of its surface functional
groups. For this reason it is useful to consider the principal kinds of surface
functional groups commonly found in soil clays.
The plane of oxygen atoms bounding a
tetrahedral silica sheet in a layer silicate is called a siloxane surface. This
plane is characterized by a distorted hexagonal (i.e., trigonal) symmetry
among its constituent oxygen atoms that is produced when the underlying
tetrahedra rotate to fit their apexes to contact points on the octahedral
sheet. Further accommodation of the tetrahedra to the octahedral sheet is
achieved through the tilting of their bases so that the silicon-oxygen bonds
arc directed toward the contact points instead of lying normal to the basal
THE SILOXANE DITRIGONAL CAVITY.

14

THE SURFACE CHEMISTRY OF SOilS

plane of the mineral. As a result of this adjustment, one of the basal


oxygen atoms in each tetrahedron is raised about 0.02 nm above the other
two and the siloxane surface becomes corrugated.
The functional group associated with the siloxane surface is the ditrigonal cavity formed by six corner-sharing silica tetrahedra (Fig. 1.7). This
cavity has a diameter of about 0.26 nm and is bordered by six sets of
lone-pair electron orbitals emanating from the surrounding ring of oxygen
atoms. These structural features qualify the ditrigonal cavity as a Lewis
base, which is any molecular unit that uses a doubly occupied electron
orbital in initiating a chemical reaction.P
The reactivity of the siloxane ditrigonal cavity depends on the nature of
the electronic charge distribution in the phyllosilicate structure. If there are
no isomorphic cation substitutions to create deficits of positive charge in
the underlying layer, the ditrigonal cavity functions as a very soft Lewis
base (electron donorjl' and is likely to complex only neutral dipolar
molecules, such as water molecules. The complexes formed are not very
stable, an example being the easily reversed entrapment of a water
molecule having one of its hydroxyl groups directed into the cavity along a
normal to the siloxane surface.!" If isomorphic substitution of AIH by
Fe2+ or Mg2+ occurs in the octahedral sheet, the resulting excess negative
charge can distribute itself principally over the 10 surface oxygen atoms of

Figure 1.7. The siloxane ditrigonal cavity in a distorted tetrahedral sheet of a


phyllosilicate. (Original computer-drawn figure courtesy of C. T. Johnston)

THE REACTIVE SOLID SURFACES IN SOILS

15

the four silica tetrahedra that are associated through their apexes with a
single octahedron in the layer. This distribution of negative charge
enhances the Lewis base character of the ditrigonal cavity and makes it
possible to form complexes with cations as well as with dipolar molecules.
An outer-sphere surface complex of this type with a Ca2+ cation is
illustrated in Fig. 1.8. In this example, which is the familiar two-layer
hydrate of Ca-montmorillonite, two opposing ditrigonal cavities in separate siloxane surfaces complex a Ca2+ cation solvated by six water
molecules in octahedral coordination.
On the other hand, if isomorphic substitution of Si4+ by AIH occurs in
the tetrahedral sheet, the excess negative charge can distribute itself
primarily over just the three surface oxygen atoms of one tetrahedron, and
much stronger complexes with cations and dipolar molecules become
possible because of this localization of charge. In particular, inner-sphere
surface complexes, such as the one illustrated with a K+ cation in Fig. 1.8,
now are more likely. This complex requires the coordination of the
potassium ion with 12 oxygen atoms bordering two opposing ditrigonal
cavities. The layer charge in illitic micas and in vermiculites is large enough
(Table 1.3) to allow each ditrigonal cavity in the basal planes of these
minerals to complex a K+ cation. Moreover, the ionic diameter of K+
(Table 1.1) is almost precisely equal to that of a cavity. This combination
of charge distribution and stereochemical factors gives the K-mica and
K-vermiculite surface complexes their well-known stability in soils. 15
The range of Lewis base character possible for the siloxane ditrigonal
cavity can be demonstrated most directly through infrared spectroscopic
studies of complexes involving hydrogen bonds. The stretching frequencies
of hydrogen-bonded NH, OH, and OD groups decrease uniformly as the
strength of the bond increases. Take, for example, the NH group in NHt,
or in an alkylammonium ion (e.g., C2H5NHt) or the OH group in H 20
Figure 1.8. Surface complexes between metal cations and siloxane ditrigonal
cavities on 2: 1 phyllosilicates, shown in exploded view.

INNER-SPHERE SURFACE COMPLEX:


K+ ON VERMICULITE

OUTER-SPHERE SURFACE COMPLEX:


CoIH 20l: + ON MONTMORILLONITE

THE SURFACE CHEMISTRY OF SOilS

16

hydrogen-bonded to a siloxane surface oxygen atom: the greater the


localization of negative charge on the oxygen atom, the stronger the
hydrogen bond. Accordingly, the strongest hydrogen bonds should form
with oxygen atoms near sites of isomorphic substitution in the tetrahedral
sheet and the weakest bonds should involve oxygen atoms that share no
excess negative charge from isomorphic substitution. Hydrogen bonds with
surface oxygen atoms near sites of isomorphic substitution in the
octahedral sheet should be of intermediate strength. These predictions
imply that the stretching frequencies of hydrogen-bonded NH or OH
groups are lower for vermiculite than for montmorillonite complexes, and
this trend is exactly what is observed experimentally in/infrared spectroscopic studies. 16
The most abundant and most reactive
surface functional group in soil clays is the hydroxyl group exposed on
the outer periphery of a mineral. This kind of OH group is found on
phyllosilicates, on amorphous silicate minerals, and on metal oxides,
oxyhydroxides, and hydroxides. In general, for a given mineral, more than
one kind of surface OH group can be distinguished on the basis of
stereochemical reasoning, and these different groups have properties (e.g.,
their infrared absorption spectra) that set them apart from OH groups
inside the bulk mineral structure.
These general characteristics of the inorganic surface hydroxyl group can
be illustrated quite well with goethite, whose molecular structure is shown
in Fig. 1.2. The surface of goethite consists primarily of exposed planes
that lie parallel to the crystallographic c axis and perpendicular to either
the a or the b axis. The surface OH groups on these planes are shown in
Fig. 1.9. Three types of OH groups are found on the plane perpendicular
to the crystallographic a axis; they are denoted A, B, and C in Fig. 1.9.
THE INORGANIC HYDROXYL GROUP.

Figure 1.9. Surface hydroxyl groups on goethite: singly (A-type), triply (B-type),
and doubly (C-type) coordinated to Fe (III) , along with Lewis acid site hydroxyls.
The drawing on the right shows an inner-sphere surface complex with HPO~- at the
A-type hydroxyl group. The dashed lines indicate hydrogen bonds.
SURFACE HYDROXYLS

LEWIS
ACID SITE

GOETHITE SURFACE HYDROXYLS


AND LEWIS ACID SITE

INNER-SPHERE SURFACE COMPLEX:


HPO~- ON GOETHITE

THE REACTIVE SOLID SURFACES IN SOILS

17

The type A hydroxyl group is a former oxygen ion coordinated to one Fe3+
cation in the bulk structure that has become protonated upon exposure as a
surface group. The type C hydroxyl group is formed in the same way,
except that it is coordinated to two Fe 3 + cations. The type B hydroxyl
group is a hydroxyl group coordinated to three Fe 3 + cations that has
become exposed on a surface. The infrared absorption spectrum of these
three surface OH groups is different from the spectrum of the bulk
structural OH groups.'?
By contrast, there are only type C hydroxyl groups on the plane
perpendicular to the crystallographic b axis, and these are always coordinated to an Fe3+ cation with an accompanying water molecule. This
arrangement, in which Fe(III) H 20 acts as a Lewis acid site, is shown in
Fig. 1.9. (A Lewis acid is any molecular unit that uses an empty electron
orbital in initiating a reaction. 13 In the present example, the Lewis acid is
the Fe 3 + cation.) The type A hydroxyl group can be protonated to form a
Lewis acid site and then be exchanged to allow the formation of an
inner-sphere complex with the oxyanion HP0~- (Fig. 1.9), which consists
of HPO~- bound through oxygen ions to two adjacent Fe3+ cations. Both
the OH in the o-phosphate unit and the oxygen ions coordinated to the
Fe3+ cations are hydrogen-bonded to the goethite surface. The size and
configuration of the o-phosphate unit are especially compatible with the
grooved structure of the goethite surface, thus providing a stereochemical
enhancement to the stability of the inner-sphere complex. Inner-sphere
complexes also can form through the exchange of protonated OH groups
and other oxyanions.l"
Surface hydroxyl groups coordinated to pairs of AI3+ cations appear on
the plane perpendicular to the crystallographic c axis in gibbsite (Fig. 1.2).
This basal plane makes up most of the surface of the mineral, but it appears
that the hydroxyl groups bound to Lewis acid sites on the edge plane
perpendicular to the basal plane are more reactive.l? These Lewis acid
sites comprise an AI3+ cation coordinated to a single water molecule.
The presence of hydroxyl groups coordinated to either one or two metal
cations is a common feature of the surfaces of oxide, oxyhydroxide,
hydroxide, and silicate minerals in soils. The phyllosilicates in particular
expose singly coordinated OH groups on the edge surfaces created when
crystallites are broken apart. These edge-surface hydroxyls are illustrated
in Fig. 1.10 for kaolinite. At the edge of the octahedral sheet AI(III)'H20,
which is a Lewis acid site, is found. The hydroxyl group associated with the
site can form an inner-sphere surface complex with a proton at low pH
values or with an hydroxide anion at high pH values. The water molecule
hound to the AI3+ cation also can be expected to be replaced by an
hydroxide anion at higher pH values. At the edge of the tetrahedral sheet,
hydroxyl groups are singly coordinated to Si4 + cations. Because of the
greater valence of silicon, these OH groups tend to complex only hydroxide anions. as opposed to the OH groups coordinated to Al(III), which

THE SURFACE CHEMISTRY OF SOilS

18

complex both H+ and OH-. The two types of edge-surface hydroxyl


groups, which are often distinguished by the names aluminol and silanol,
also differ in their reactivity toward oxyanions such as HPO~-, in that
SiOH does not form inner-sphere surface complexes comparable with those
shown in Fig. 1.9. 18 However, inner-sphere and outer-sphere surface
complexes with metal cations are possible for either kind of hydroxyl
group.i" An outer-sphere complex between an ionized edge-surface H 20
and Na+ is illustrated in Fig. 1.10.
A variety of functional groups are
present in the organic compounds that polymerize to form the humic
substances in soil clays (Fig. 1.4), and it is expected that some of these
functional groups ultimately reside on the interfaces between solid soil
organic matter and the fluid phases in soils. The more prominent organic
surface functional groups in soils are given in Table 1.5. !,erhaps the most
significant of these in well-oxidized soil organic matter are the carboxyl,
carbonylvand phenylhydroxyl(phenolic 0H)grQups,21 The stabilities of
co'ffiplexes'between
key' groups and protons range from weak
(uncharged carbonyl) to very strong (phenolic OH, which does not ionize
until about pH 9). 1.:.~~s. a~pep.trum of surf~c.:J:llncti2Ila!.~E()1.lP}~~1J.c~hitris
likely, and superimposed on this intrinsic variability is that created by the
wloe range of stereochemical and charge distribution characteristics possibI~ .in a heterogeneous . orgaIlic.!!llltrix. For this reason it is entirely
conceivable thatthe properties of organic surface functional groups are not
well defined, but instead can be characterized onlY.1:>Y ranEes of values.
'''featllreth<l;L.
This "smea!iIlg out" of chemicaCJ~~hiYiQtJi:iij"imPQ~ia~t
aist~~g~ishe.s_.()ri~nJc .s~,~!~~~j~tiQ11~lg~()1l,p's from functionli" . . gl"o~ps
bound to small organic molecules.
ORGANIC SURFACE FUNCTIONAL GROUPS.

these'

Figure 1.10. Surface hydroxyl groups (shaded) on kaolinite. Besides the OH


groups on the basal plane, there are aluminol groups, associated with Lewis acid
sites, and silanol groups protruding from the edge surface. The right side of the
figure shows an outer-sphere surface complex between an ionized H 2 0 and Na +, as
well as complexes between the silanol groups and OH- (i.e., proton dissociation).

H 20

LEWIS
ACID SITE
ALUMINOL

SILANOLS

KAOIINI TF SURFACE HYDROXYLS

IONIZED SILANOLS

OUTER-SPHERE SURFACE COMPI F X:


N<l(fV))~

ON K/IOI INI T I

THE REACTIVE SOLID SURFACES IN SOILS

19

Table 1.5. The organic surface functional groups in soil


clays
Functional
group

Structural
formula

Relative stability of
proton complexes

o
-Carboxyl
,,,..-Carbonyl
Amino
Imidazole
---Phenylhydroxyl
Sulfhydryl

II

-C-oH

Weak to moderate

-C-

Very weak
Strong
Strong
Very strong
Very strong

o
II

-NHz
Ring NH
Ring OH
-SH

o
Sulfonic

II

-S-OH

Very weak

II
o

1.3. SOLID SURFACES IN NATURAL SOILS

Soils in nature are heterogeneous assemblies of solid materials that are


altered continuously through processes initiated by biological, geological,
and hydrological agents. This transience, which is characteristic of the
open boundaries between the lithosphere, atmosphere, and hydrosphere,
brings with it a pattern of..biQlQgic.ally meqiJ!ti'g..c:;h~mi~~U~::i!!.f9.E!!l:~t.~9.!l.~
of the solid phases i!1~qil that is known aspedo(;~ertlfcal w~q!~~r,ing. The
surface chemistry of soils is conditioned on pedochemical weathering. The
dissolution and precipitation of minerals and the oxidation and synthesis of
organic matter necessarily alter the nature and reactivity of surface
functional groups. For example, the weathering of a smectite to a kaolinite
and ultimately to gibbsite and colloidal silica certainly transforms the
surface chemistry of a soil clay fraction. In this case, the sequential progress
ofpedochemical weathering can be epitomized as an evolution of the suite of
surface functional groups from preeminence of the siloxane ditrigonal
cavity to preeminence of the inorganic hydroxyl group. This kind of
pedochemical impact on surface functional groups forms part of the
temporal context for the structural descriptions and classifications introduced in Sec. 1.1 and 1.2. In the present section, three important examples
of this temporal context are discussed to illustrate its effect on surface
functional groups.
The structural similarity in illitic mica. vermiculite. and smectite minerals
was noted in Sec. 1.1 (Table 1.3). The basal planes of these minerals are

20

THE SURFACE CHEMISTRY OF SOILS

siloxane surfaces that permit stacking of the 2: 1 layers along the crystallographic c axis while not promoting strong bonding between the individual
layers in a stack. On the basis of these relationships alone, a close
association of the three phyllosilicate groups in soils can be expected, and
this association is in fact reported often in pedogenesis studies.F These
studies indicate that the inner-sphere surface complex between the siloxane ditrigonal cavity and the potassium ion (Fig. 1.8) plays a critical role in
the weathering processes that relate the 2: 1 phyllosilicates.P Soil conditions that favor the formation of this complex (e.g., low concentrations of
competing monovalent cations in the soil solution, low water content,
moderate temperatures) also tend to stabilize illitic mica against transformation to vermiculite or smectite. Vermiculite in soil clay differs from
illitic mica principally in the properties of its surface complexes (there is
usually very little difference in layer charge). These complexes ordinarily
are mixtures of inner-sphere complexes involving K+ and outer-sphere
complexes involving Na +, Ca2+, or Mg 2 + . When the fraction of innersphere surface complexes is near 0.5, the likelihood of finding some
combination of illitic mica and vermiculite layers in contact becomes very
great. The stability of the inner-sphere surface complex with K+ depends
significantly on the orientation of its neighboring OH group in the
octahedral sheet (Fig. 1.8). If an outer-sphere complex forms on one
siloxane surface of a phyllosilicate layer, the structural OH group nearest
the complex tends to point more toward it than if an inner-sphere complex
were there because a complexed solvated cation is farther away from the
bottom of the ditrigonal cavity. This effect allows the structural OH group
nearest the ditrigonal cavity in the siloxane surface on the opposite side of
the phyllosilicate layer to point more away from that cavity since there is
now less proton-proton repulsion. Because of this orientation of the OH
group, inner-sphere complexes on the opposite siloxane surface are
especially stable and therefore quite likely to form. The net result of this
sequence of interactions is stacks of illitic mica and vermiculite layers with
the two minerals in regular alternation.P
A similar phenomenon can occur with smectite to produce this kind of
regular interstratification of illitic mica layers with, for example, montmorillonite layers in soil clays. However, the fact that the layer charge on
smectite originates primarily in the octahedral sheet and is significantly
smaller than that on illitic mica tends to decrease the stability of smectite
inner-sphere complexes and to reduce the chance that a strictly regular
stacking of the phyllosilicate layers will occur. Since a decrease in layer
charge and isomorphic substitutions in the octahedral sheet also can take
place in vermiculite, phyllosilicate mixtures in soil clays commonly exhibit
only partial regularity or even random ordering in the stacked layers,
especially if the concentration of K + in the soil solution is low.?" If, in
addition to structural irregularities caused by isomorphic substitutions in
these phyllosilicate layers, there is also a relatively high concentration of
protons in the soil solution. the formation of inner-sphere surface complexes with K I is inhihited completely and a transformation of the layer

THE REACTIVE SOLID SURFACES IN SOILS

21

type from 2: 1 to 1: 1 becomes favorable. Under these conditions, it is


possible to have random interstratification of smectite and kaolinite
minerals as an intermediary in the weathering process instead of stacked
mixtures containing illitic mica layers."
The siloxane surface of a 2: 1 phyllosilicate bears a negative layer charge
regardless of the composition of the soil solution. This property makes it
possible for the ditrigonal cavities in the surface to react collectively with
positively charged polymers to form interlayer surface coatings. In a
similar fashion, inorganic surfaces bearing hydroxyl groups can react with
negatively charged polymers in the soil solution and become coated.
Pedochemical weathering brings about the dissolution of minerals that
release elements in high oxidation states into the aqueous phase of soil,
and it is these elements that hydrolyze readily to form hydroxy-polymers.
The most important of these polymers contain Al(lII) and Si(IV). Aluminum hydroxy-polymers are metastable dissolved species whose formation
and complexation by the interlayer siloxane surfaces of vermiculite and
smectite are favored by pH values below 6.0, by low concentrations of
organic compounds, and by frequently varying water content." Their
presence on an interlayer siloxane surface can be interpreted as an
intermediate step in the formation of dioctahedral chlorite (Fig. 1.3).
Conversely, the gradual stripping of the hydroxide interlayer in chlorite
observed in some soil clays can be viewed as a weathering regression
toward vermiculite or smectite. In this case, interstratification of the
chlorite mineral with the vermiculite or smectite is expected as an
intermediate step in the process.i" It is possible for iron and even
magnesium hydroxy-polymers to be complexed in a similar fashion by
siloxane ditrigonal cavities; however, the observed incidence of these
complexes in soil clays is very low.25 In the case of silica polymers, the
more probable surface complexes are those with the exposed hydroxyl
groups on kaolinite, gibbsite, and, to some extent, goethite.Ir"
There is evidence also for surface complexation of organic polymers in
soil clays.27 The mechanisms through which these complexes form are
numerous, but there is general agreement that both inner-sphere and
outer-sphere complexes can form with each class of surface functional
group described in Sec. 1.2. Stereochemistry plays a much more important
role in surface organic complexes than it does in complexes with inorganic
polymers, and therefore fewer generalizations are possible since the
structure of dissolved organic matter in soils is so poorly understood. The
configurations and bonding mechanisms in surface organic complexes are
considered in Chap. 4.
The three examples of the effects of pedochemical weathering on the
surface structures in soil clays just described illustrate the complexity of the
reactive solid materials in natural soils. To these examples can be added
many others, including the formation of iron oxyhydroxide or calcium
carbonate coatings on the external surfaces (as opposed to interlayer
surfaces) of phyllosilicates, the development of thick envelopes of colloidal
organic matter on uggrcgutes of metal oxides and aluminosilicates, lind the

22

THE SURFACE CHEMISTRY OF SOILS

growth of two-dimensional solid solutions of trace metal oxides on the


periphery of manganese oxide nodules. The complications introduced by
these features of the surfaces in soil clays cannot be dismissed in a study of
the surface chemistry of soils, but neither should they be accorded a
dominant role without clear experimental data in support of that position.
A soil solution that contains at least 1 millimole of protons per cubic
meter (pH = 6) along with much dissolved organic matter but few metal
cations can be expected to alter the surface structure of phyllosilicates as
illustrated in Fig. 1.11. The two principal effects are incorporation of
hydroxy-polymers into the interlayer region and onto edge surfaces and
development of iron oxyhydroxide and organic coatings on the external
surfaces of stacks of layers which themselves may contain mixtures of
different phyllosilicates.Fr" Similar effects occur on the surface of oxide
and carbonate minerals. Under these conditions, the significant chemistry
of the surface functional groups in the clay fraction is principally that of
the groups exposed on the hydroxy-polymers and on the external coatings instead of the groups on the underlying matrix. Even if these new
polymeric surface groups are not numerous, they still can be very
important in the adsorption reactions of metal cations and oxyanions found
in low concentations in the soil solution.j" On the other hand, if soil pH

Figure 1.11. Some effects of pedochemical weathering on phyllosilicate surfaces.


Weathering produces interlayer hydroxy-polymers, interstratification, and external-surface organic and inorganic polymer coatings on smectite. On kaolinite,
organic and iron oxide coatings are produced by weathering. (After Jenne 29 )

UNWEATHERED SMECTITE LAYERS

UNWEATHERED KAOLINITE
LAYERS

hydroxy pol ymer

mica layer

SMECTITE-POLYMER COMPLEXES

oroanlc coat Ino

KAOLINITE-POLYMER
COMPLEXES

23

THE REACTIVE SOLID SURFACES IN SOilS

values are not low, if the soil solution is rich in metal cations but not in
organic compounds, and if leaching has not been intensive, the surface
chemistry of the clay fraction may be related directly to that of the
uncoated solid matrix, especially if the reactions of major elements, e.g.,
Ca(II) or Cl( - 1), are under investigation.
1.4. SPECIFIC SURFACE AREA

The specific surface area. of a soil clay sample is the combined surface area
9tal~t~epllrticles in the sample as determinedby sOIIle experi.fil~~tal
techniqu: and expresse~pe~llni.!!!1ass of t~~ ~"ll.~ple. Thus the SI units of
surrace~area" are 'square .meters per kilogram. As its definition
implies, s.pecific surface area is afl operationalconcept. Tlle.Dumeric'!.Lya1.l1~
f.?J!!ld for a"~~Y~!1 s9n~~1"~i'."g5m~Jm~..QJJ.JYbiQu~~en.talm~lhg2j !!1~~q.
There are two principal reasons for this very important characteristic.
First, the properties of the solid surfaces in a soil clay often can be altered
during the preparation of the clay for a surface area measurement. For
example, it rn~.YJ)~~.ece~l!IY~!S;L,9J:Y,Jh~~c'1ay thoroughly and maintain it
under vacuum, or treatlll:ents~ith,~h,eJI.l.tca,lr~llg~n,..t~may be required for
purposes of sample ~standardizatlon..Both of tl)._~:techniql1.~s~~!l:.<ll.!~Ltbe.
shape and,si?;(;f, gf scil, particlesalld,y'()i<:l,p;:l~~es. Second, if a surface
reaction is involved in the measurement of specific surface area, the data
obtained reflect only the ceh,5l~llct~ri!ics.()f the ~llrfllce. fu!wtiQllalgrQ.MPS
in~af!lciR~~3.~~!~~r~~!i.<?~. Both stereoch~mical and charge ?istribunon effects can lImIt the kinds of surface functional groups accessible to a
particular molecule used as a surface probe, and therefore ~.S.h.lKOli9,al
measurem~.I1t.ofspecifif,sllrf~,lCe areain general provid.es.()Illy inf()~~~at~()n
ibouTi:he' solid....,.._..,.,.-._",
surfaces that were
reactive under the conditions of the
, ,."'.'.

specific .

meas"llrement.~

<, "

"

".""

..- _

,'

_ . ,

0.E~!ational n!iture of specific surfac~are.a precludes any interpretation of its numerical values in an absolute geometric sense. There is
no specific surface area of a soil clay, but only specific surface areas, each
determined with some surface chemical application in mind. If the extent
of sample alteration produced by required pretreatment is large, then the
soundest use of the numerical results from a given method is simply a
comparison of values for different soil clays prepared under standardized
circumstances. If a chemical reaction is the basis for the measurement,
then the results are meaningful only if applied to molecules similar to the
probe molecule reacting with surfaces similar to those in the measured
sample. It is this operational context that underlies the discussion of
experimental methods to follow.

. The

The principal physical methods for measuring specific surface areas of soil clays arc electron microscopy and X-ray diffraction. In both techniques. the objective is to determine the shapes and
dimensions of representative soil particles and then to calculate the
PHYSICAL METHODS.

--

24

THE SURFACE CHEMISTRY OF SOILS

crystallographic specific surface area, So, with the equation


particle surface area
(1.3)
(particle mass density) (particle volume)
where the particle surface area and volume are evaluated with geometric
formulas based on particle dimensions and shape. The particle mass
density can be measured in a separate experiment or can be obtained from
published data.
Consider as an example of the use of Eq. 1.3 a clay sample containing !t~~~hi!~ ..sr,yl!tals (paralle~epipeds) wi~h ~he dimensions 160 x 26 x
6.5 nm. Since the mass density of goethite IS 4.37 x 103 kg . m- 3, the
crystallographic specific surface area of these crystals is
So =

So =

2[(160 x 26) + (160 x 6.5) + (26 x 6.5)]


160 x 26 x 6.5 x 4370
9.09 x 10- 5 m3nm- 1kg- 1 = 9.09 x 104 mZkg- 1

= 90.9 mZg- 1

In a similar fashion, the crystallographic specific surface area of hexagonal


!faQHnjl~,l::Iy.stals (Fig. 1.11) of thickness 43 nm and with a distance of
1<P nm separating directly opposing edges on the hexagonal faces is
So =

(2 x face area) + (6 x edge area)


face area x thickness x density

1.881 x 10 6 nrrr'
-------::;---;;:--------::;-------;;:(3.723 x 107 nm 3)(2.615 x 103 kg . m 3)
-1.93 x 1O-5m3nm- Ikg- 1 = 1.93 x 104m Zkg- I

= 19.3mZg- 1

where the mass density of kaolinite is taken as 2.615 x 103 kg . m-3.


The use of l?~?!i.:'h~4I].~~~~.~~~:.~i!X.~!~ in these two examples is
pr~9ic'!te,q9:n. .~.~e,.. !!.s.~}!.!!11?!!~~,.t~~~.!~~~h~..T..~~!..~2~!tio~ o{.1b e smL
.par!ifle. is the s~,~~."~~.!h~!.?f ,the,.~~7.Ini~~llXp\};r.e,~ES!!!l~11,..Q!UYll.icll.!h.e
density.me,~sure~e,I!t.~.~e,~e,In!1de,. In soil clays this assumption does not
apply if isomorphic substitutions have occured to any extent, as they do,
for example, in the 2: 1 phyllosilicates (Table 1.3). For these kinds of
minerals, it is possible to use data on the chemical composition directly in
the calculation of the specific surface area if X-ray crystallographic
information is also available. This alternate method of calculation entails
computations of the relative molecular mass of a unit cell, based on the
structural formula for the mineral, and of the area of a quasicrystal formed
by stacking the phyllosilicate layers along the crystallographic c-axis
(Fig. 1.11).30
The computation can be illustrated with a montmorillonite whose
structural formula is NaO.66 Si8 (A I3.o5 Feo.z9Mgo.66)Ozo(OHk The relative
molecular mass of the unit cell is calculated by multiplying the relative
molecular mass of each clement by its stoichiometric coefficient and
summing over ull such products to ontain M, ."" 742.4 for the total. Note

THE REACTIVE SOLID SURFACES IN SOilS

25

that this result depends both on the pattern of isomorphic substitution and
on the type of exchangeable cation. The unit cell dimensions for the
mineral are" a = 0.517 nm, b = 0.895 nm, and c = 0.95 nm. The surface
area per unit cell is then 2ab = 0.9254 nnr' per cell if the edge surfaces are
neglected. This neglect of the edge surfaces is justified by the fact that
montmorillonite particles typically are plates whose lateral dimensions
(about 100 nm) greatly exceed their thickness (1 to 6 nm), so that the edge
surface area contributes only a few per cent to the total. The crystallographic specific surface area of an Avogadro number of unit cells of montmorillonite now can be calculated:
NA
per cell x ---=-=
So = surface area
----'0-relative mole mass of cell

0.9254 nm 2 x 6.022 x 1023


742.4 g

7.51

1020 nm 2g- 1

= 7.51 x 105 nr'kg" = 751 m2g- 1

where N A = 6.022 X 1023 is the Avogadro constant. This well-known


result applies to platy particles whose thickness equals that of the unit cell.
If instead there are n layers stacked along the crystallographic c-axis to
form a quasicrystal, So is calculated by dividing 751 m2g- 1 by n, since each
pairing of stacked layers places two siloxane surfaces inside the quasicrystal. The procedure illustrated here applies equally well to all the phyllosilicates listed in Table 1.3.
Besides the problems of surface alteration that may arise in connection
with sample preparation for electron microscopy and X-ray diffraction
studies, there is also the difficulty of obtaining sufficient data to characterize the broad range of particle shapes and dimensions found in natural
soil clay. The magnitude of effort implied by this heterogeneity limits the
scope of the physical methods of determining specific surface area to
establishing reference values for particular groups of minerals and to
calibrating other methods of surface area measurement for selected clay
samples. Faced with an array of soil particles like those portrayed in
Fig. 1.11 one cannot expect physical methods to enjoy wide application as
a routine approach to the measurement of surface area.
POSiTIVE, ADSORPIIQN,M,m:tQI>.S. The measurement of specific surface
area with a positive adsorption method requires the satisfaction of three
basic experimental conditions:

1. There must be a chemical reaction between the surfaces of the clay


particles and a molecular unit that results in a pronounced accumulation
of the molecular unit on the surfaces, i.e., positive adsorption. The
reaction may be between a degassed clay sample and a compound in
the gas phase or between a clay in aqueous suspension and a dissolved
solute.

26

THE SURFACE CHEMISTRY OF SOILS

2. The mass of adsorbate per unit mass of clay corresponding to one layer
of adsorbate on the clay surfaces must be determined. The term
adsorbate refers to the accumulating molecular unit mentioned in
condition 1.
3. The 1'~,~~!I?:g,,~J:~a2Lt~e.~~~.?rp~!~,~!JE.?~?I.':lyer coveE~~~ ,mus.t_be
Qs:t~.nuil,l,ed. The packin? a~eais the aInount?fsurf~ce~rea alloted to
each adsorbed mOiecurari.init:~--,,"...... ,,~, . , . " " - ,_... "".
s,

When conditions 1 to 3 have been met, the adsorption specific surface area,
in the SI units of square meters per kilogram, can be calculated with the
equation

sm

= XMm N Am
a x 10- 15

(1.4)

where X m is kilograms of adsorbate per kilogram of sample at monolayer


coverage, M, is the relative molecular mass of the absorbate, and am is its
packing area in square nanometers.
The adsorptive (i.e. adsorbable compound) used to determine specific
surface area can be chosen, on the basis of its molecular properties, to
react selectively with the particular surface functional groups whose areal
distribution is of interest. Often, rather weakly interacting, nonselective
adsorptives are used when the objective is to measure as much of the
exposed surface area as possible. For example, nitrogen gas is a commonly
used adsorptive in surface area measurements because it interacts weakly
with a broad array of surface functional groups and therefore permits the
determination of exposed surface area in many different kinds of clay. The
principal limitation on the use of this adsorptive is stereochemical, since
the relatively large van der Waals radius of the N 2 molecule prevents it
from interacting with surface functional groups occluded in very small void
spaces.
The packing area of an adsorbate molecule in the gas phase often has a
value close to that predicted by assuming that the molecule forms an ideal
monolayer, i.e., one with hexagonal close packing at the density of its bulk
liquid phase:
(1.5)
where p is the mass density of the bulk liquid in kilograms per cubic meter
and a~ is expressed in square nanometers. The fact that calculated values
of am are close to a~ suggests that the packing area is to some degree the
intrinsic property it is supposed to be. However, as shown in Table 1.6,
packing areas determined through calibration studies on well-characterized
solid surfaces vary for the same adsorbate, especially in the case of water
vapor.
The value of the monolayer parameter, X m , in Eq. 1.4 is determined
universally with an adsorption isotherm equation. If it is known from
experiment that the adsorption forms no more than a single molecular
layer on the surfaces of interest, X m is then the x intercept of the line

27

THE REACTIVE SOLID SURFACES IN SOilS

Table 1.6. Packing areas of gas-phase adsorbates used in specific surface area
measurements
Gas-phase
adsorbate

T,K

Nitrogen
Argon
Krypton
Ethylene glycol
(1,2-ethanediol)
Ethylene glycol monoethyl ether
(2-ethoxy-ethanol)
Water

aid
m-

nrrr' am, nm 2

Range of
observed
am, nm 2

78
77
78

0.162
0.138
0.152

0.162
0.167
0.202

0.13-0.20
0.13-0.18
0.17-0.22

293

0.224

0.332

0.230-0.332

293
293

0.323
0.105

0.523
0.106

0.396-0.600
0.075-0.195

obtained by fitting adsorption data to a linear form of the Langmuir


equationr"
(1.6)
where q is the mass of adsorbate per unit mass of sample, K is an empirical
constant, and K d is the distribution coefficient, the ratio of q to the
corresponding adsorptive concentration (aqueous solution) or pressure
(gas phase) at equilibrium. The use of Eq. 1.6 to determine X m is illustrated
in Fig. 1.12. The derivation of Eq. 1.6 is discussed in Chap. 4. For
gas-phase adsorptives, it is common practice to equilibrate the sample with
the adsorbate at a pressure high enough to make the approximation
q = X m valid. This "single-point" method of determining X m has the

Figure 1.12. The calculation of the monolayer parameter, X m , for Ncetylpyridinium bromide (CPB) adsorption (Langmuir plot) and water vapor
adsorption (BET plot) by soils. 33 35

BET PLOT
H20 Adsorption
Otago podzol

xm=1/(23.4+ 1.74) =
0.0398kg H20/kg soil
oL-.

L--_ _--l.--L_ _--J

os

1.0

q(mol kcr I)

I.~

oL--_---l_ _--l._ _--L_ _--J

0.10

0.30

0.40

28

THE SURFACE CHEMISTRY OF SOILS

distinct advantage of convenience for routine measurement and is widely


used with either ethylene glycol or ethylene glycol monoethyl ether as the
adsorbate.F However, the combined uncertainties in the determination of
X m by a single measurement and the present lack of extensive calibration of
the packing area restrict this approach to use only for a qualitative
comparison of specific surface areas in soil clays.
If the adsorbate is in the gas phase and tends to form many layers on the
surfaces of soil clay particles, as do water vapor and the inert gases listed in
Table 1.6, then X m can be calculated as the reciprocal of the sum of slope
and y intercept of the line obtained by fitting adsorption data to a linear
form of the Brunauer-Emmett-Teller equationi"

P/ Po
q(1 - p/Po)

1
(C - 1) P
xmC + xmC Po

(1.7)

where P is the pressure of the adsorbate, Po is its pressure at saturation


(i.e., in equilibrium with a bulk liquid phase of the adsorptive), and C is an
empirical constant. A graph of the left side of Eq. 1.7 versus the relative
pressure, p / Po, is usually a straight line for adsorption data obtained in the
domain 0.05 <: p/Po <: 0.30. The use of Eq. 1.7 to calculate X m is also
illustrated in Fig. 1.12. The physical model on which the BET equation is
based requires that the probabilities for evaporation and condensation of
the adsorbate from each molecular layer formed, as well as the energies
associated with these two processes, be independent of the amount
adsorbed, s" This assumption turns out to be valid approximately in the
range of relative pressures indicated above.
A combination of nitrogen gas and water vapor is frequently used to
measure the adsorption specific surface areas of soil clays containing
principally phyllosilicates.P Because of its size and weak interaction
characteristics, the N2 molecule is adsorbed only on the external surfaces
of phyllosilicate quasicrystals. The water molecule, on the other hand, can
penetrate a smectite or vermiculite quasicrystal to enter interlayer regions
containing inner-sphere complexes between metal cations and the siloxane
ditrigonal cavities, provided that these complexes are intrinsically less
stable than outer-sphere surface complexes involving the solvated metal
cations would be. The intercalation of water molecules to solvate the
complexed metal cations is favored when the latter have high ionic
potential (the ratio of valence to ionic radius) and can strongly attract and
orient dipolar molecules. Therefore, cations such as Li +, Na +, Mg2+, and
Ca2+, when complexed by phyllosilicates, are associated with interlayer hydration, whereas K+, Rb ", and Cs" tend not to be. X-ray diffraction data
show that, when interlayer hydration occurs, X m determined as indicated in
Fig. 1.12 actually corresponds to a single layer of intercalated water
molecules." Therefore, X m for water vapor adsorption refers to monolayer
coverage of the external surfaces plus one half of the accessible internal
surfaces of the quasicrystals. This physical interpretation of X m for water
vapor, as well as, that of X m measured by nitrogen gas adsorption, permits

THE REACTIVE SOLID SURFACES IN SOilS

29

the calculation of the accessible internal plus external adsorption specific


areas of illitic mica, vermiculite, and smectite minerals: If SN represents
the specific surface area calculated with Eq. 1.4 from nitrogen adsorption
data and Sw is that calculated from water vapor adsorption data, then the
total adsorption specific surface area is
(1.8)

according to the interpretation of X m values given previously.


The adsorption behavior of N-cetylpyridenium bromide (CPB), an
organic molecule comprising a cetyl hydrocarbon chain substituted onto a
pyridine ring that weakly complexes a bromide ion (C16H33CsHsN+B;),
on phyllosilicates is similar to that of water vapor. CPH can be adsorbed
from dilute aqueous solution by phyllosilicates and used to determine their
specific surface areas." The adsorption follows the Langmuir equation,
with X m calculated according to Eq. 1.6 (Fig. 1.12). On external surfaces,
monolayers of CPB molecules are adsorbed in head-to-tail pairs, with one
pyridine-ring head bound to the surface and the cetyl-chain tail oriented
perpendicularly. The packing area is 0.27 nrrr'. This same arrangement is
found in the interlayer region, but here both pyridine rings in an adsorbed
pair of CPB molecules are in contact with siloxane surfaces. Therefore, X m
again corresponds to coverage of the external surfaces plus one half of the
accessible internal surfaces and the adsorption specific surface area can be
calculated with Eq. 1.8.
Typical results of specific surface area determinations on phyllosilicates
by nitrogen gas/water vapor or nitrogen gas/CPB adsorption are listed in
Table 1.7. For Mg-vermiculite and Na-montmorillonite, the measured
adsorption specific surface area is close to that calculated from the unit cell
dimensions and structural formula. For illitic mica, the area is about 14
per cent of the ideal crystallographic value, indicating that this mineral
forms particles containing about seven phyllosilicate layers that cannot be
penetrated by water vapor or CPB.
Although the positive adsorption methods offer the advantage of
convenient determination in heterogeneous samples, they suffer from the
uncertainty involved in the calibration of the packing area (which usually
must be done by comparison with results from a physical method using
reference clay materials) and from the fact that the monolayer parameter is
model-dependent through Eqs. 1.6 and 1.7. It must also be remembered
that the specific surface area determined by positive adsorption is ultimately a function of the reaction between surface functional groups and
some probe molecule. If the experimental conditions of the reaction are
close to those under which the surface behavior of a sample is of interest,
then this estimate of specific surface area has surface chemical relevance.
Negative adsorption refers to the
phenomenon in which a charged solid surface confronts an ion of like
charge in an aqueous suspension and the ion is repelled from the surface by
NEGATIVE

ADSORPTION

METHODS.

THE SURFACE CHEMISTRY OF SOilS

30

Table 1.7. Specific surface areas of phyllosilicates determined by nitrogen, water

vapor, or N-cetyl pyridinium bromide adsorption


Sm (nitrogen +

water),
104 m 2kg- 1
Kaolinite (Naform, Peerless)"
Kaolinite (Bath,
South Carolina?
Illitic mica (Naform, Illinois)"
Illitic mica (Naform, Fithian, Illinois)"
Vermiculite (Mg-form,
llano, Texas)"
Vermiculite (Kenya)"
Montmorillonite (Naform, Wyoming)"
Montmorillonite (Naform, Wyoming)'
a

CPB),
104 m 2kg- 1

1.88

1.8

1.5

10.2
9.3

9.6

0.31
<0.1

71.2

3.3

84.7

1.4

72.6

80.0

A. G. Keenan et a!., J. Phys. Colloid Chern. 55: 1462 (1951).

D.
C H.
d H.
e R.
b

1.86

Sm (nitrogen

J. Greenland and J. P. Quirk, J. Soil Sci. 15: 178 (1964).


D. Orchiston, Soil Sci. 78:463 (1954).
van Olphen, Proc. Int. Clay Cant 1969 1:649 (1969).
W. Mooney et a!., J. Am. Chern. Soc. 4: 1367 (1952).

D. J. Greenland and C.J.B. Mott, in The Chemistry of Soil Constituents (D. J. Greenland and M.H.B.
Hayes, eds.), p.321. Wiley, Chichester, U.K., 1978.

coulomb forces. The coulomb repulsion produces a region in the aqueous


solution near the surface that is relatively depleted of the ion and a
corresponding region, far from the surface, that is relatively enriched in the
IOn.
This effect can be observed experimentally in the following way. An
aqueous solution containing a chosen ion i is poured into two identical
chambers separated by a membrane permeable to water and dissolved
solutes but not to suspended solids (i.e., a dialysis membrane). The moles
of charge contributed by ion i to one of these chambers is IZ i IeOi V, where
Z, is the valence of the ion, COi is its concentration, and V is the volume of
the chamber. After m; kilograms of a solid are suspended in one of the
chambers and equilibrium with respect to the transport of ion i is established across the membrane, the concentration of ion i in the chamber not
containing the suspended solid will rise to c, if the surfaces of the solid repel
ion i. This increase in concentration is associated quantitatively with a
region of depIction of ion i in the chamber containing the suspended solid

THE REACTIVE SOLID SURFACES IN SOILS

31

through the definition

IZ i ICi V ex

IZilcoY

= !ZilciV - '---'-'----'-------'-----'ms

(1.9)

where Vex is called the exclusion volume. The numerator on the right side
of Eq. 1.9 is equal to the number of moles of charge of ion i sent into the
chamber not containing the suspended solid. The exclusion volume thus
represents the volume per unit mass of suspended solid depleted of ion i in
the chamber containing the suspension.
The development of a negative adsorption method for measuring specific
surface area is based on the additional definitiorr"
(1.10)
where dex(Ci) is the exclusion distance, a function of the concentration, c..
The parameter SE is the combined solid surface area per unit mass from
which ion i is repelled in an aqueous suspension. If this surface is also the
entirety of that bathed by the aqueous solution, then SE as measured by
the combination of Eqs. 1.9 and 1.10 is the full specific surface area of the
suspended solid. The parameter d ex represents the mean distance over
which the ion i is depleted near the surfaces of the suspended solid. It is
evaluated conventionally as a function of c, with the help of the diffuse
double layer (DDL) theory of an ion swarm near a charged planar surface
in aqueous suspension. 37
The model calculation of dex(c;) is presented here for the important
special case of a negatively charged solid suspended in a 1: 1 electrolyte
solution. According to the DDL theory,37 the surface charge density
neutralized by a swarm of electrolyte ions is

O'a = -{2EoDRTc[exp( -Fl/Ja/RT) + exp(Fl/Ja/RT) - 2]}1/2 (1.11)


where O'a is the surface charge density (coulombs per square meter), EO is
the permittivity of vacuum, D is the dielectric constant of liquid water, R is
the molar gas constant, T is the absolute temperature, C is the same as c,
(subscript suppressed), F is the Faraday constant, and l/Ja is the electric
potential (volts) at the plane where the diffuse ion swarm comes into
contact with the solid. (The derivation of Eq. 1.11 is discussed in Chap. 5).
Often the condition - Fl/Ja/RT ~ 1 is met and Eq. 1.11 can be approximated with the expression

O'a = -(2EoDRTc)1/2exp( -Fl/Ja/2R T)

(1.12)

Equation 1.12 and the standard DDL relationship.F

O'a

= -EoD(dl/J/dx)x=a

(1.13)

then lead to the differential equation

dl/J

-, =
(X

(2RTC/F.f1[)

1/ 2

cxp(-Fl/J/2RT)

(X = 8)

(1.14)

32

THE SURFACE CHEMISTRY OF SOILS

where i3 is the distance between two planes: that where l/J = where u" is evaluated. The solution of Eq. 1.14 is

exp(Fl/J,,/2RT)

F(2c/ eoDRT)1/2 i3/2

00

and that
(1.15)

The definition of dex(c) in DDL theory is38

dex(c) =

L'' [1 -

(1.16)

exp(Fl/J(x)/RT)]dx

The right side of Eq. 1.16 is the relative probability that a univalent anion
will not be found at a point x near a negatively charged planar surface,
integrated over all x values from i3 outward. Thus the mean exclusion
distance is equal to the probability that an ion is excluded from the region
between x and x + dx, summed over all such regions. The integral in
Eq. 1.16 can be calculated with the help of a transformation based on
Eqs. 1.11 and 1.13:

dex(c)

= (0 1 - exp(Fl/J/RT)
)"'3

dl/J/ dx

{O

(1 - eY)dy
~ )Y3 [c(e- Y + eY - 2)]1/2
1

dl/J

= -1-

jj3C

iO ey/

dy

Y.

= r;;: [1 - exp(y,,/2)]

v f3c

2
= - - i3

(1.17)

jj3C

where

f3 == 2F 2/e oDRT = 1.084

1016 m . mol- 1

(T

298.15 K)

and y == Fl/J/RT. The last step in the calculation is made with the help of
Eq. 1.15.
The introduction of Eq. 1.17 into Eq. 1.10 produces the DDL model
equation for the exclusion volume in a 1: 1 electrolyte:
(1.18)
Equation 1.18 predicts that a graph of measured values of Vex versus the
function C- 1/2 will be a straight line with a slope proportional to the
exclusion specific surface area, SE' This behavior is illustrated in Fig. 1.13
for montmorillonite suspended in NaCI.:w A least-squares line through the
data points in the graph has a slope equal to 0.3046 mol l / 2 dml/2kg - I.

33

THE REACTIVE SOLID SURFACES IN SOILS

No-MONTMORILLONITE IN NoCI
10kg CLAY/ m3 SUSPENSION

)(

>0)
Vex = 0.5524

+ 0.3046c- 1/ 2

r 2 = 0.968

0'---------4-------'-------'--------'

30

10

40

Figure 1.13. A plot of the exclusion volume against electrolyte concentration


according to Eq. 1.18 for Na-montmorillonite suspended in NaC1. 39

Therefore,
SE

=~

x (slope/2)

(1.084)1/2
10- 3 / 2 m 3 / 2
1/2
1/2
8
=
x 10 m
molx ----;;-;",.-2
dm 3 / 2
x 0.3046 mo1 1/2 dm 3 / 2kg- 1
= 5.01 x

lOS rrr'kg ?

= 501 m 2g- 1

The surface area of the montmorillonite sample that repels chloride ions
amounts to about 500 m 2g- 1 under the conditions of the experiments.
Table 1.8 lists specific surface area values for illitic micas as determined
by nitrogen gas adsorption and by negative chloride adsorption.t" The
specific surface areas calculated from N2 gas adsorption with the help of
Eq. 1.7 show no particular trend with type of exchangeable cation. The
mean value of SN, 11.2 0.5 x 104 m2kg-1, suggests that the mineral
forms particles containing seven phyllosilicate layers, as indicated previously. The external surfaces of these particles are expected to repel
anions, and therefore the specific surface area determined by negative
chloride adsorption should also be around 105 rrr'kg" '. As shown
in Table 1.8. however, the values of Sf:. obtained with Eq. 1.18, are
always less than SN and decrease sharply with increasing radius of the

34

THE SURFACE CHEMISTRY OF SOilS

Table 1.8. Specific surface areas of illitic micas and montmorillonites determined
by Nz gas adsorption (SN) and chloride exclusion (SE)40
Exchangeable
cation
Li+
Na+
K+
NHt

Rb+
Cs+

Illitic mica, 104 mZkg- 1

Montmorillonite, 104 mZkg- 1

SN

SE

SN

SE

11.6
10.9
11.3
11.8
10.5
11.2

8.0
7.0
3.5
2.2
1.0
0.0

6.6
4.6
6.4
5.9

65.0
56.0
43.6
25.6

14.6

15.6

exchangeable cation. The cause of this trend is thought to be surface


complex formation between, for example, siloxane ditrigonal cavities and
the exchangeable cations. 40 Surface complex formation results in the
creation of an external surface that is electrically neutral wherever the
complexes occur. Therefore, the extent of the surface that can repel
chloride anions is reduced below that probed by N2 molecules and the
specific surface area estimated with Eq. 1.18 is less than SN' The fact that
this difference increases with cation size can be attributed to the higher
tendency of larger cations to form inner-sphere surface complexes because
of their stereochemistry and ionic potential. 41
Table 1.8 also compares specific areas for montmorillonite as determined by N z gas adsorption and negative chloride adsorption.t" In this
case, the specific surface areas obtained with Eq. 1.18 are less than the
crystallographic value of 72.5 x 104 rrr'kg" but are generally larger than
SN' The variation in SE with type of cation can be understood in terms
of both surface complexation and quasicrystal formation. Small-angle
neutron scattering experiments on montmorillonite suspensions suggest
that Li- and Na-saturated clays are likely to remain as single layers.F If
this is true, the reduction in exclusion specific surface area below the
crystallographic specific surface area for Li- and Na-montmorillonite indicated in Table 1.8 must reflect a corresponding reduction, because of
surface complexes, in the amount of surface that can repel chloride anions.
Surface complexation evidently decreases the charged surface area by 10
and 23 per cent, respectively, for Li- and Na-clay. On the other hand,
K-montmorillonite can form quasicrystals containing two phyllosilicate
layers in suspension and Cs-montmorillonite can form quasicrystals containing three layers.F If a fraction f of the total number of phyllosilicate
layers in a suspension form quasicrystals containing n layers, then Eq. 1.10
must be replaced by the expressiorr"
Vex =

f f(
1-

1-

~) lSodex(C) + ~ Sod

( 1.19)

THE REACTIVE SOLID SURFACES IN SOilS

35

where So is the crystallographic specific surface area and d is the fixed


distance between opposing internal surfaces in a quasicrystal. The derivation of Eq. 1.19 assumes no surface complexation and no chloride anions
inside the quasicrystals. It is evident from Eq. 1.19 that the specific surface
area determined by negative adsorption is related to the crystallographic
value according to the equation
(1.20)
This expression can reproduce the values of SE for K- and Cs-rnontmorillonite in Table 1.8 if / is given the values 0.80 and 1.2, respectively.
The impossibly high value of/for Cs-montmorillonite emphasizes that part
of the difference between SE and So for Cs-montmorillonite and the other
clay samples must come from surface complexation of the exchangeable
cations.
The differences between the nitrogen gas and chloride exclusion specific surface area values for montmorillonite in Table 1.8 bring into relief
one of the more important advantages of the negative adsorption method.
Since this method requires that an aqueous solution come in contact with
the solid surface, the specific surface area measured should reflect the
structure of the solid material in natural soil more than, for example, the
gas adsorption methods. Evidently the degree of aggregation of montmorillonite (except for Cs-montmorillonite) in the degassed state is quite
different from what it is in aqueous suspension, with the result that the
surface available to repel chloride anions is much larger than that which
can adsorb N z moleclues, despite the important effect of surface complexation. With illitic mica minerals, however, it appears that the state of
aggregation of the samples does not change a great deal when they are
brought into suspension from the dry state, but surface complexes do exert
a profound effect on the specific surface area determined by chloride
exclusion.
1.5. SURFACE CHARGE DENSITY

It is well established that some of the surface functional groups in soil clays
bear electric charge, the sign and magnitude of which depend on the
composition of the soil solution and the structure of the solid phase to
which the functional groups are bound. The siloxane ditrigonal cavity, for
example, often bears a more or less localized negative charge produced by
isomorphic cation substitutions, and the s~~~_~x~2~~Lg~9,Y1?.'!JlJJ.~i!~
either positive or negative charge depending on thl1.pJ:!":~!~~,9iJh~~2!1
solution. Because the ~e.ac~ive solid surfaces in soils are h.e~er()Ae.!l_C:l:).~~.,.!!!e
c:Oiicepf of surface charge density for .th.~.~ is.plu.~.~!i_~~i.c, not,~?n.o..typ~<::;
The several kinds of surface charge density in a soil clay are discussed in
Chap. 3. In the present section. attention is devoted to introducing surface
charge density as an ~"era,~()nal c(!nc:C'I" through a description of some of

THE SURFACE CHEMISTRY OF SOILS

36

the methods used to measure it and are interpretation of these methods


from the point of view of surface functional group chemistry.
The surface density of intrinsic
charge is the number of coulombs per square meter borne by surface
functional groups either because of isomorphic substitutions in soil minerals or because of proton association and dissociation reactions. The
intrinsic surface charge density, O"in, thus refers to permanent structural
charge and to the net charge produced by proton-selective functional
groups (e.g., aluminol groups) on both inorganic and organic surfaces.
This charge density can be measured conveniently by the Schofield
method.P which consists of reacting soil clay with an electrolyte solution at
a fixed pH value, removing excess electrolyte to retain only adsorbed
cations and anions, and determining the moles of cation and anion charge
adsorbed by a unit mass of the clay. The intrinsic surface charge density is
the product of the Faraday constant and the difference between adsorbed
cationic and anionic moles of charge per unit mass of clay, divided by the
specific surface area of the clay:
INTRINSIC

SURFACE

CHARGE

DENSITY.

- F(q+ - q_)

(1.21)

where q+ and q_ refer to the moles of adsorbed cation and anion charge,
respectively. Note that O"in can be either positive or negative.
The chemical interpretation of O"in measured by the Schofield method
depends sensitively on the type and concentration of probe electrolyte
used. If these properties are chosen so that the cation in the reacting
electrolyte neutralizes precisely the exposed functional group charge
associated with isomorphic substitutions and dissociated hydroxyls and so
that the anion neutralizes only the exposed protonated functional groups,
then q + and q _ will have optimal magnitude for the chosen pH value and
O"in will be truly an intrinsic surface charge density. On the other hand, if
the cation in the probe electrolyte is not able to displace all of the native
adsorbed cations in, e.g., inner-sphere surface complexes, or if the anion
cannot displace all of the native anions bound to protonated functional
groups, or if either ion does not form only neutral surface complexes in the
soil clay, then O"in will differ from its optimal value.
Thus the intrinsic surface charge density viewed operationally can exhibit
different values for the same soil clay at a given pH. The optimal value of
O"in represents the difference between the largest quantities of positive and
negative charge achievable intrinsically by exposed surface functional
groups. Nonoptimal values of O"in fall into the broad spectrum of possible
differences between the varying amounts of positive and negative charge
that can be brought to soil clay surfaces by cations and anions of varying
adsorption characteristics. If they are positive, these nonoptimal values of
q +. and q. usually are termed cation and anion exchange capacities.
respectively. The reflect only the reactivities of chosen probe ions with

THE REACTIVE SOLID SURFACES IN SOILS

37

surface functional groups under prescribed conditions. However, if these


probe ions and experimental conditions are similar to those in the natural
soil clay, the value of Uin they produce can be of practical utility even if it is
not optimal. 44
The surface density of permanent
structural charge, uo, is the coulombs per square meter borne by surface
functional groups that are charged because of isomorphic substitutions
in soil minerals. The sign of Uo is nearly always negative in moderately
weathered soils and is associated with the charge on siloxane ditrigonal
cavities near sites of isomorphic substitution in 2: 1 phyllosilicates. The
magnitude of Uo thus can be estimated from chemical composition and
X-ray crystallographic data, following a procedure similar to that described
in Sec. 1.4 for calculating the crystallographic specific surface area of a
Na-montmorillonite. In that example, chemical composition data indicated
0.66 moles of negative charge, created by the substitution of Mg(II) for
Al (III) in the octahedral sheet, per mole of unit cells. Since the basal plane
dimensions of the unit cell were a = 0.517 nm and b = 0.895 nm, it
follows that, for Uo in coulombs per square meter,
STRUCTURAL SURFACE CHARGE DENSITY.

qF)
5
= (2N
(10

18

Uo

Aab

(- 0.66 mol cmol- 1)(9.64870 x 104 C'mol;I)(10 18 nmZm- Z)


2(6.022 x lO Z3 mol- 1)(0.517 nm)(0.895 nm)
= -0.114 Cm- z
(1.22)

where Bq is the charge deficit, F is the Faraday constant, and N A is the


Avogadro constant. The same method can be applied to calculate Uo for
any 2: 1 phyllosilicate.
The magnitude Uo can also be estimated from the properties of the
complexes formed between siloxane ditrigonal cavities and N-alkylammonium cations, CnHz n + 1NHj .45 The procedure consists of reacting
a Na-saturated soil clay with a series of N-alkylammonium chloride
solutions at 65C, washing the reacted clay with ethanol to remove excess
electrolyte, and determining the basal plane [d(001)] spacing in the clay
after drying it under vacuum.:" The basal plane spacing is then plotted
against n c , the number of carbon atoms in the N-alkylammonium cation
used in an adsorption experiment, with the entire series of experiments
included in the range 1 < n; < 20. When a monolayer of complexed
alkylammonium cations lies in the interlayer region, the basal plane
spacing is 1.36 nm and when a bilayer is present, it is 1.77 nm. Besides these
conformations of the adsorbed organic cations, one observes pseudotrilayers, in which each alkyl chain in the bilayer kinks toward the basal plane
opposite the one it is lying upon, and paraffin-type structures, in which
each alkyl chain makes a nonzero angle with the basal plane complexing
the NH~ cation."

38

THE SURFACE CHEMISTRY OF SOILS

When N-alkylammonium cations form the 1.36-nm monolayer structure, they lie flat between opposing basal planes and each cation covers a
van der Waals area equal to (0.057n c + 0.14) nm z. Since the cation is
univalent, this area is associated with one proton charge. As the value of n;
increases, the area required by a cation increases. At some point, the area
required becomes larger than the area per electron charge on a basal plane
and the monolayer structure is no longer stable. The 1.77-nm bilayer
structure then becomes the favored one because the area requirement of
the adsorbed organic cation can be met independently on each opposing
basal plane in the interlayer region (Fig. 1.14). At the monolayer-bilayer
transition, the area per adsorbed cation just equals the area per electron
charge on a basal plane. Therefore, at the transition point,
ab
0.057n c + 0.14 = x
x =

ab
0.057n c + 0.14

(1.23)

where x is the layer charge of the phyllosilicate, defined in Sec. 1.1. With
the values of x indicated in Table 1.3 and a typical value of 0.46 nm z for
the product ab, the monolayer-bilayer transition can be expected for
4 <: n; :5 14 in smectite and 2 <: n c <: 4 in vermiculite. The layer charge of
illitic mica is usually too large for the monolayer-bilayer transition to be
observed at any n.:
Figure 1.14 illustrates the monolayer-bilayer transition for a clay fraction
deep in the subsoil of a Spanish Vertisol."? As is commonly found, the
Figure 1.14. Basal plane spacings of N-alkylammonium complexes with a Vertisol
clay fraction and the corresponding layer charge distributiorr."
bilayer

monolayer

Layer Charge
Distribution

o
~

~ 50

~2.0

E
.5
o

9 1.5
-01

....'

15

a..

SOIL CLAY
VERTISOL

1.0
n

u
cr
w

....L...L..........,
20

0,6

LAYER CHARGE

0.7

THE REACTIVE SOLID SURFACES IN SOILS

39

transition does not take place at a single value of n.; instead it begins at
n c = 11 and ends at n c = 14. This gradual transition reflects the fact that
the layer charge in the clay does not have a single value, i.e., there is layer
charge heterogeneity. In this example, the layer charge distribution can be
estimated with the help of Table 1.9, which is based on Eq. 1.23 (with
ab = 0.465 nm 2 ) and on the theory of X-ray diffraction by randomly
interstratified phyllosilicates. 48 The basal plane spacings in Fig. 1.14 are
1.36,1.48,1.60, and 1.77 nm for n c = 11,12,13, and 14, respectively. The
data in Table 1.9 show that basal spacings of 1.36 nm at nc = 11 and
1.48 nm at n c = 12 correspond to about 29 per cent of the interlayer
regions having layer charges between 0.58 and 0.64. Similarly, 1.6 nm at
n c = 13 and 1.48 nm at n c = 12 correspond to 49 - 29 = 20 per cent
of the interlayer regions having layer charges between 0.56 and 0.58, and
1.77 nm at n c = 14 and 1.6 nm at n c = 13 correspond to 51 per cent of the
interlayer regions with layer charges between 0.52 and 0.56. This distribution of layer charge is plotted in Fig. 1.14. Similar kinds of charge
partitioning involving combinations of X-ray diffraction data with theoretical analysis are possible for vermiculitic and mixed-phyllosilicate soil
clays."?
DENSITY. The surface density of net proton
is defined by an expression analogous to Eq. 1.21:

PROTON SURFACE CHARGE

charge,

UH,

(1.24)
where qH is the moles of complexed proton charge and qOH is the moles of
complexed hydroxyl charge on proton-selective surface functional groups

Table 1.9. Relations between carbon number (n c ) and layer charge (x), and between
basal plane spacing [d(OOl)] and bilayer
fraction (p) for
N-alkylammoniumsmectite complexes'"

6
7
8
9

10
11

12
13
14
15

d(OOl), nm

1.00
0.88
0.80
0.74
0.68
0.64
0.58
0.56
0.52
0.50

1.36
1.40
1.45
1.50
1.55
1.60
1.65
1.70
1.73
1.77

0.00
0.13
0.24
0.33
0.40
0.49
0.58
0.70
0.80
1.00

"Bused on Elj, 1,23. with 11/1 ~ 0.46~ nm'. corrected


for u purticlc diumctcr of IlKI

11111,

40

THE SURFACE CHEMISTRY OF SOILS

per unit mass of soil clay. The quantity qOH is equal formally to the moles
of charge on ionized, proton-selective functional groups per unit mass of
soil clay. The numerator in Eq. 1.24 applies only to surfaces bearing
functional groups whose charge is intrinsically pH-dependent, i.e., inorganic hydroxyl groups and most organic functional groups.
The maximum values of qH/S and qOH/S for oxide, hydrous oxide, and
hydroxide minerals can be estimated from crystallographic data if supporting information concerning the potential reactivity of the surface hydroxyl
groups is available. For example, although the plane perpendicular to the
crystallographic a axis on the surface of goethite contains four OH groups
per unit cell (one type A, two type B, and one type C), only the type A
group is believed to react with protons to form OHt groups.P Since the
unit cell dimensions of goethite are a = 0.465 nm, b = 1.002 nm, and
c = 0.304 nm", there should be one reactive OH group per 0.305 nrrr',
corresponding to a maximum qH/S of 5.45 x 10- 6 mol.m "? when the
plane is fully protonated. Given that 80 per cent of the goethite surface
is made up from this plane, the maximum value of qH/S is 4.36 x
10- 6 mOlcm- Z The maximum value of qOH/S for goethite can be estimated
similarly after making the assumption that all of the type A hydroxyl
groups plus the water molecules bound to the Lewis acid sites (Sec. 1.2)
can ionize at high pH. Full dissociation of the water molecules, which lie on
the plane perpendicular to the crystallographic b axis that makes up the
rest of the goethite surface, yields one OH- per 0.141 nrrr', or 1.18 x
10- 5 mOlcm- z. The resulting maximum value for qOH/S is 6.72 x
10- 6 mOlcm- Z Crystallographic estimates of maximum qH/S and qOH/S
for goethite, gibbsite, and kaolinite are presented in Table 1.10. In each
calculation, the assumption was made that only aluminol groups can be
protonated, whereas aluminol, silanol, and Lewis acid water molecules can
each dissociate one proton. It was assumed also that only the edge surfaces
on gibbsite and kaolinite bear reactive hydroxyl groups. This assumption is

Table 1.10. Crystallographic estimates of the maximum values of qHISand qOHIS


for selected soil minerals

qOH/S,
p,molcm -z

Mineral
Goethite
Gibbsite
Kaolinite
a

4.4
2.8
0.35

6.7
5.6
1.0

Reactive hydroxyl groups assumed"


Type A OH and Lewis acid OHz
Edge-surface Lewis acid OH and OHz
Edge-surface silanol and aluminol,
Lewis acid OH z

Geometric disposition of the OH groups:


Goethite -one OH per 0.305 nm 2 on the plane perpendicular to a axis (80% of total surface) and
one OH 2 per 0.141 nrrr' on the planc pcrpendicular to b axis (20% of total surfacc)
Gibbsite -one OH and one OH 2 per 0.246 nm 2 on the edge surfaces (41,'1% of totul surface)
Kaolinite-s-one silanol, one alumino/, and one Olll per (U7'1 nm 2 on the edge surfaces (7,'1% of
h'lal surfacc)

THE REACTIVE SOLID SURFACES IN SOILS

41

controversial at present but not inconsistent with the results of ion


adsorption experiments.Vr'"
The most common method for measuring (qH - qOH) under arbitrary
conditions is by potentiometric titration. 50 Briefly, the method involves the
use of a glass electrode and a double-junction calomel reference electrode
in the titration cell:
glass
electrode

suspension of
solid in background
electrolyte

background
electrolyte
solution

liquid
junctions

calomel
electrode

The emf of the electrode assembly is measured while known volumes of


either acid or base are added to the suspension. These data, in turn, are
converted to proton concentrations with the help of a calibration curve
prepared from similar titration data obtained without the suspended solid
in the cell. The values of (qH - qOH) then can be calculated with the
expression
(1.25)
where CA is the molar concentration of added acid, CB is that of added
base, [H+] is the molar proton concentration derived from emf measurements, Cs is the mass of solid in one cubic decimeter of the suspension, and

cK
[OH-] = [H:]

(1.26)

where C K; is the conditional equilibrium constant for the ionization of


water in the background electrolyte solution. As it stands, Eq. 1.25
provides values of qH - qOH referred to an arbitrary zero point. Usually
these empirical values are renormalized to the value of qH - qOH at the
point of zero salt effect (PZSE), which is the pH value at which plots of
qH - qOH versus pH obtained at different ionic strengths in the background electrolyte solution meet. 50 Formally, this point of intersection is
defined by the equation

iJUH)
aI

= 0

(pH = PZSE)

(1.27)

where I is the ionic strength of the background electrolyte solution and Tis
the absolute temperature. The PZSE and other points of zero charge are
discussed in Chap. 3.
Equation 1.25 fundamentally is a mass balance expression for protons
and hydroxide ions added to a suspension of surface-reactive solids. If V is
the volume of the suspension, then (C A - [H+])V and (C B - [OH-])V
are equal, respectively, to moles of protons and moles of hydroxide ions
"bound" in some sense by the suspension constituents. The difference
between these two quantities divided by the muss of the solid material in

42

THE SURFACE CHEMISTRY OF SOILS

the suspension is taken to be qH - qOH' In order for this equality to


hold, there must be only two final states for a proton or an hydroxide ion
added to the suspension: the free ion and the ion complexed by a
functional group on a surface whose charge is pH-dependent. Therefore,
no added proton or hydroxide ion can react with dissolved constituents to
form soluble complexes, with solid phases to dissolve them, or with
surfaces whose charge is not pH-dependent. If these proton-consuming
side reactions occur, the values of qH and qOH will be overestimated in
the analysis of titration data. In practical terms, the suppression of side
reactions for H+ and OH- requires a background electrolyte whose
component ions do not form significant complexes with protons or
hydroxide ions (e.g., NaCI0 4 ) , as well as the elimination of dissolved
carbon dioxide, either by purging or by calibration procedures. The
unwanted reactions with solid phases are more difficult-perhaps even
impossible-to suppress if a natural soil clay is being titrated. Should the
soil clay contain siloxane surfaces bearing cations that can exchange with'
protons, or should it contain hydroxy polymers that consume protons
readily and dissolve, there can be little expectation of obtaining values of
qH that have surface chemical significance. 50 For this reason, the use.Qf
}J()teIltipmetric titration to measur~ ?"flITla~ bel!~~!~~t~ ,,~~~~.J.l~ions
~I!!PE~~.iPg ~~!!.~~ara~teJi~ed metal()xides,.~)'~~ousoJ{i~~.~?()rEy~.~?xides

c;>r to s\.lspeps!Qns. of purified QfgapiC::.mate.riilI.

NOTES
1. The geometric relationships among solid particle shape, area, and volume are
described in detail in L. D. Baver, W. H. Gardner, and W. R. Gardner, Soil
Physics, Chap. 1. Wiley, New York, 1972.
2. A detailed summary of this and other perturbations of the octahedral sheet is
given in R. E. Grim, Clay Mineralogy, Chap. 4. McGraw-Hill, New York,
1968.
3. L. Pauling, The Nature of the Chemical Bond, Chap. 13. Cornell University
Press, Ithaca, N.Y., 1960.
4. U. Schwertmann and R. M. Taylor, Iron oxides, in Minerals in Soil Environments (J. B. Dixon and S. B. Weed, eds.) Soil Science Society of America,
Madison, Wis., 1977. U. Schwertmann, D. G. Schulze, and E. Murad, Identification offerrihydrite in soils by dissolution kinetics, differential X-ray diffraction, and Mossbauer spectroscopy, Soil Sci. Soc. Am. J. 46: 869 (1982).
5. H. D. Megaw, Crystal Structures: A Working Approach, Chap. 13. Saunders,
Philadelphia, 1973.
6. R. M. McKenzie, Manganese oxides and hydroxides, in J. B. Dixon and
S. B. Weed, op. cit:"
7. G. Brown, A.C.D. Newman, J. H. Rayner, and A. H. Weir, The structures
and chemistry of soil clay minerals, in The Chemistry of Soil Constituents
(D. J. Greenland and M.H.B. Hayes, eds.). Wiley, Chichester, U.K., 1978.
Concerning goethite, see also R. W. Fitzpatrick and U. Schwertmann, AI
substituted goethite: An indicator of pedogenic and other weathering environments in South Africa, Geoderma 27:33; (I \1H2).

THE REACTIVE SOLID SURFACES IN SOILS

43

8. C. E. Weaver and L. D. Pollard, The Chemistry of Clay Minerals. Elsevier,


Amsterdam, 1973. The term illitic mica refers to a dioctahedral micaceous
mineral weathered in soil.
9. G. Sposito, The chemical forms of trace metals in soils, in Applied Environmental Geochemistry (I. Thornton, ed.). Academic Press, London, 1983.
10. S. Wada and K. Wada, Density and structure of allophane, Clay Minerals
12: 289 (1977). K. Wada, Mineralogical characteristics of andisols, in Soils with
Variable Charge (B.K.G. Theng, ed.). New Zealand Society of Soil Science,
Lower Hutt, N.Z., 1980.
11. W. Flaig, H.Beutelspacher, and E. Rietz, Chemical composition and physical
properties of humic substances, in Soil Components, Vol. 1: Organic Components (J. E. Gieseking, ed.). Springer-Verlag, New York, 1975.
12. A. G. Walton, Polypeptides and Protein Structure. Elsevier, New York, 1981.
13. See Chap. 3 of G. Sposito, The Thermodynamics of Soil Solutions (Clarendon
Press, Oxford, 1981) for an introduction to Lewis acids and bases.
14. R. Prost, Interactions between adsorbed water molecules and the structure of
clay minerals: Hydration mechanism of smectites, Proc. Int. Clay Conf. 1975,
p. 351 (1976).
15. A detailed discussion of the charge distribution and stereochemical aspects
of these complexes is given in V. C. Farmer and J. D. Russell, Interlayer
complexes in layer silicates, Trans. Far. Soc. 67:2737 (1971).
16. V. C. Farmer and J. D. Russell, Infrared absorption spectrometry in clay
studies, Clays and Clay Minerals 15:121 (1967). V. C. Farmer, The characterization of adsorption bonds in clays by infrared spectroscopy, Soil Sci. 112:62
(1971). H. E. Doner and M. M. Mortland, Charge location as a factor in the
dehydration of 2:1 clay minerals, Soil Sci. Soc. Am. I, 35:360 (1971).
17. R. J. Atkinson, Crystal morphology and surface reactivity of goethite. PhD
dissertation, University of Western Australia, Perth, 1969. J. D. Russell,
R. L. Parfitt, A. R. Fraser, and V. C. Farmer, Surface structures of gibbsite,
goethite, and phosphated goethite, Nature 248:220 (1974). R. L. Parfitt,
R. J. Atkinson, andR. SeC. Smart, The mechanism of phosphate fixation by
iron oxides, Soil~c:i.Soc. Am. r. 39:837 (1975). R. L. Parfitt, J. D. Russell,
and V. C. Farmer, Confirmation of the surface structures of goethite (aFeOOH) and phosphated goethite by infrared spectroscopy, l.C.S. Faraday I
72:1082 (1976).
18. A review of the molecular structure of inner-sphere surface complexes between
anions and surface OH groups is given in R. L. Parfitt, Anion adsorption by
soils and soil materials, Advan. Agron. 30: 1 (1978).
19. R. L. Parfitt. A. R. Fraser, and V. C. Farmer, Adsorption on hydrous oxides.
II. Oxalate, benzoate and phosphate on gibbsite, I, Soil Sci. 28: 289 (1977).
20. P. W. Schindler, Surface complexes at oxide-water interfaces, in Adsorption of
Inorganics at Solid-Liquid Interfaces (M. A. Anderson and A. J. Rubin, eds.).
Ann Arbor Science, Ann Arbor, Mich., 1981. J. A. Davis, R. O. James, and
J. O. Leckie, Surface ionization and complexation at the oxide/water interface. I. Computation of electrical double layer properties in simple electrolytes, J, Colloid Interface Sci. 63: 480 (1978).
21. F. J. Stevenson, Humus Chemistry, Chap. 9. Wiley, New York, 1982.
22. R. C. Reynolds, Interstratified clay minerals, in Crystal Structures of Clay
Minerals and Their X-ray Identification (G. W. Brindley and G. Brown, eds.).
Mineralogical Society. London. 19HO.

44

THE SURFACE CHEMISTRY OF SOilS

23. K. Norrish, Factors in the weathering of mica to vermiculite, Proc. Int. Clay
Conf. 1972, p. 417 (1973).
24. B. L. Sawhney, Interstratification in layer silicates, in J. B. Dixon and
S. B. Weed, op. cit.4
25. C. I. Rich, Hydroxy interlayers in expansible layer silicates, Clays and Clay
Minerals 16: 15 (1968). P. H. Hsu, Aluminum hydroxides and oxyhydroxides,
in J. B. Dixon and S. B. Weed, op. cit.4
26. B. D. Mitchell, Oxides and hydrous oxides of silicon, in Soil Components,
Vol. 2: Inorganic Components (1. E. Gieseking, ed.). Springer-Verlag, New
York, 1975.
27. B.K.G. Theng, Clay-polymer interactions: Summary and perspectives, Clays
and Clay Minerals 30: 1 (1982). K. R. Tate and B.K.G. Theng, Organic matter and its interactions with inorganic soil constituents, in B.K.G. Theng,
op. cit. 1o
28. E.A.C. Follett, W. J. McHardy, B. D. Mitchell, and B.F.L. Smith, Chemical
dissolution techniques in the study of soil clays, Clay Minerals 6: 23 (1965).
E.A.C. Follett, The retention of amorphous, colloidal "ferric hydroxide" by
kaolinites, J. Soil Sci. 16: 334 (1965). A. W. Fordham and K. Norrish, Electron
microprobe and electron microscope studies of soil clay particles, Aust. J. Soil
Res. 17:283 (1979).
29. E. A. Jenne, Trace element sorption by sediments and soils-Sites and
processes, in Molybdenum in the Environment (W. Chappel and K. Petersen,
eds.). Dekker, New York, 1976.
30. J. P. Quirk and L.A.G. Aylmore, Domains and quasi-crystalline regions in
clay systems, Soil Sci. Soc. Am. J. 35: 652 (1971).
31. S. Brunaer, L. E. Copeland, and D. L. Kantro, The Langmuir and BET
theories, in The Solid-Gas Interface (E. A. Flood, ed.), Vol. 1. Dekker, New
York, 1967. Equation 1.7 is derived under the assumption that an infinite
number of layers build up on the absorbing surface. If the number of layers is
finite, a more general expression results, but it cannot be distinguished from
Eq. 1.7 when plotted as in Fig. 1.12 unless the number of layers is fewer than
three.
32. D. L. Carter, M. D. Heilman, and C. L. Gonzalez, Ethylene glycol monoethyl
ether for determining surface area of silicate minerals, Soil Sci. 100: 356 (1965).
M. D. Heilman, D. L. Carter and C. L. Gonzalez, The ethylene glycol
monoethyl ether (EGME) technique for determining soil-surface area, Soil
Sci. 100:409 (1965). L. J. Cihacek and J. M. Bremner, A simplified ethylene
glycol monoethyl ether procedure for assessment of soil surface area, Soil Sci.
Soc. Am. J. 43: 821 (1979). A "single-point" method involving the adsorption
of water vapor by a Ca-saturated soil in equilibrium with a relative humidity
of 20 per cent (e.g., a saturated solution of CaBrz) has been suggested by
J. P. Quirk, cited in note 33. For critical studies of these single-point methods,
see I. M. Eltantawy and P. W. Arnold, Reappraisal of the ethylene glycol
monoethyl ether (EGME) method for surface area estimations of clays, J. Soil
Sci. 24:232 (1973), and Ethylene glycol sorption by homoionic montmorillonites, J. Soil Sci. 25:99 (1974).
33. H. D. Orchiston, Adsorption of water vapor. I: Soils at 25 C, Soil Sci. 76: 453
(1953). J. P. Quirk, Significance of surface areas calculated from water vapor
sorption isotherms by use of the BET equation, Soil Sci. 80: 423 (1955).
34. R. W. Mooney. A. G. Keenan. and L. A. Wood. Adsorption of water vapor
by montmorillonite. I: IIcut of desorption lind application of RET theory.

THE REACTIVE SOLID SURFACES IN SOILS

45

J. Am. Chem. Soc. 74: 1367 (1952). II: Effect of exchangeable ions and lattice
swelling as measured by X-ray diffraction, J. Am. Chem. Soc. 74: 1371 (1952).
35. D. J. Greenland and J. P. Quirk, Surface areas of soil colloids, in Transactions
of Comm. IV and V. International Society of Soil Science, Palmerston North,
N.Z., 1962. Determination of surface areas by adsorption of cetyl pyridinium
bromide from aqueous solution, J. Phys. Chem. 67: 2886 (1963). Determination of the total specific surface areas of soils by adsorption of cetyl pyridinium
bromide, J. Soil Sci. 15: 178 (1964).
36. R. K. Schofield, Calculation of surface areas from measurements of negative
adsorption, Nature 160: 408 (1947). R. K. Schofield and O. Talibudeen,
Measurement of the internal surface by negative adsorption, Disc. Faraday Soc.
3:51 (1948). H. J. van den Hul and J. Lyklema, Determination of specific
surface areas of dispersed materials by negative adsorption, J. Colloid Interface
Sci. 23: 500 (1967).
37. G. Sposito, The Thermodynamics of Soil Solutions, Chap. 6. Clarendon Press,
Oxford, 1981.
38. G. H. Bolt, Soil Chemistry. B: Physico-Chemical Models, Chap. 7. Elsevier,
Amsterdam, 1979.
39. G. H. Bolt and B. P. Warkentin, The negative adsorption of anions by clay
suspensions, Kolloid-Z. 156: 41 (1958).
40. D. G. Edwards, A. M. Posner, and J. P. Quirk, Repulsion of chloride ions by
negatively charged clay surfaces, I, II, and III, Trans. Faraday Soc. 61: 2808
(1965).
41. P. J. Sullivan, The principle of hard and soft acids and bases as applied to
exchangeable cation selectivity in soils, Soil Sci. 124:117 (1977).
42. D. J. Cebula, R. K. Thomas, and J. W. White, Small angle neutron scattering
from dilute aqueous dispersions of clay, J.C.S. Faraday I, 76:314 (1980).
43. R. K. Schofield, Effect of pH on electric charges carried by clay particles,
J. Soil Sci. 1: 1 (1949). B. van Raij and M. Peech, Electrochemical properties of
some Oxisols and Alfisols of the tropics, Soil Sci. Soc. Am. J. 36: 587 (1972).
D. J. Greenland, Determination of pH dependent charges of clays using
caesium chloride and X-ray fluorescence spectrography, Trans. 10th Int. Congr.
Soil Sci. (Moscow) 11:278 (1974). D. J. Greenland and C.J.B. Mott, Surfaces
of soil particles, in D. J. Greenland and M.H.B. Hayes, op. cit.'
44. This important application of nonoptimal values of ain is discussed in
R. L. Parfitt, Chemical properties of variable charge soils, in B.K.G. Theng,
op. cit. 1O
45. G. Lagaly and A. Weiss, Determination of the layer charge in mica-type layer
silicates, Proc. Int. Clay Conf. 1969, p. 61 (1969). G. Lagaly and A. Weiss,
The layer charge of smectic layer silicates, Proc. Int. Clay Conf. 1975, p. 157
(1976). G. Lagaly, M. Fernadez Gonzalez, and A. Weiss, Problems in layercharge determination of montmorillonites, Clay Minerals 11: 173 (1976).
46. G. Ruehlicke and E. E. Kohler, A simplified procedure for determining layer
charge by the N-alkylammonium method, Clay Minerals 16: 305 (1981).
47. J. L. Perez Rodriguez, A. Weiss, and G. Lagaly, A natural clay organic
complex from Andalusian black earth, Clays and Clay Minerals 25: 743 (1977).
48. G. Lagaly, Characterization of clays by organic compounds, Clay Minerals
16:1 (1981).
49. G. Lagaly, The "layer charge" of regular interstratified 2: I clay minerals,
Clays and Clay Minerals 27: I (1979). Layer charge heterogeneity in vermiculites, Clays and Clay Minerals 30: 215 (19H2).

46

THE SURFACE CHEMISTRY OF SOilS

50. For good working discussions of potentiometric titration methods, see G. H.


Bolt, Determination of the charge density of silica soils, J. Phys. Chern.
61:1166 (1957), and D. E. Yates and T. W. Healy, Titanium dioxideelectrolyte interface. 2. Surface charge (titration) studies, J. C.S. Faraday I
76: 9 (1980). A critical discussion of the uses of potentiometric titration to
measure O'H for soil clays is given in J. C. Parker, L. W. Zelazny, S. Sampath,
and W. G. Harris, Critical evaluation of the extension of zero point of charge
(ZPC) theory to soil systems, Soil Sci. Soc. Am. J. 43: 668 (1979).
FOR FURTHER READING

J. B. Dixon and S. B. Weed, Minerals in Soil Environments. Soil Science Society of


America, Madison, Wis., 1977. Chapters 4 through 12, 16, and 18 of this
comprehensive treatise may be consulted for detailed accounts of the structural
chemistry of the solid phase in soils.
J. E. Gieseking, Soil Components, Vol. 2: Inorganic Components. SpringerVerlag, New York, 1975. Chapters 1 through 8 of this encyclopedic reference
provide perhaps the best advanced discussions available on the structural chemistry
of phyllosilicates.
D. J. Greenland and M.H.B. Hayes, The Chemistry of Soil Constituents. Wiley,
Chichester, U.K., 1978. The first four chapters in this outstanding anthology of soil
chemistry may be consulted as background for the topics discussed in the present
chapter. Every aspect of soil mineralogy considered in detail in the first two entries
of this reading list is described in Greenland and Hayes with equal authority but
more briefly.
S. J. Gregg and K.S.W. Sing, Adsorption, Surface Area and Porosity. Academic
Press, London, 1982. The first two chapters of this well known monograph present
a thorough discussion of the concept of the packing area and the measurement of
specific surface area by positive adsorption methods.
S. J. Gregg and K.S.W. Sing, The adsorption of gases on porous solids, Surface and
Colloid Science 9: 231 (1976). A shortened version of Adsorption, Surface Area and
Porosity for the reader who wishes a brief review of the concept of packing area and
the BET method.
\
G. D. Parfitt and K.S.W. Sing, Characterization of Powder Surfaces. Academic
Press, London, 1976. Chapter 2 offers an especially good review of the physical and
chemical methods for characterizing surface hydroxyl groups.
B.K.G. Theng, Soils with Variable Charge. New Zealand Society of Soil Science,
Soil Bureau, Department of Scientific and Industrial Research, Lower Hutt, N.Z.,
1980. The first 10 chapters of this fine compendium survey the mineralogical and
surface chemical properties of soils whose reactive solid surfaces are populated
principally by the hydroxyl group.

2
THE STRUCTURE OF WATER NEAR
CLAY MINERAL SURFACES

2.1. LIQUID WATER

The most important surface reactions in soils occur when liquid water is
the fluid phase in contact with the particles of the clay fraction. An understanding of these reactions requires not only information about the structural chemistry of the solid phases, the subject of Chap. 1, but also an
appreciation of what effects the contiguous liquid phase may have on
surface functional group behavior. As a first approximation, one can state
that these effects should be the same as those observed for functional group
reactivity in aqueous solutions containing small molecules. This conjecture
can be correct for dilute soil clay suspensions, but it can be quite wrong for
soil clays enveloped only by a thin film of water because the solid surfaces
could perturb the water molecules enough to alter their configuration from
what exists in the bulk liquid phase. The altered water structure, in turn,
could exhibit solvent properties different from those of the bulk liquid and
therefore affect surface functional group reactivity in a different manner.
Any inference concerning the effects of a possibly altered molecular
structure of water near the solid surfaces in soil clays must proceed from an
acquaintance with the structure of liquid water in bulk and in aqueous
electrolyte solutions. In this section, the current picture of the molecular
arrangement in bulk water is reviewed. In Sec. 2.2, the same is done for
aqueous solutions of inorganic electrolytes. These summaries are followed
by discussions of the structure of water near the surfaces of phyllosilicates
and the effect of these surfaces on the solvent properties of the water
molecule.
The characteristic feature that distinguishes liquids from solids, and in
particular liquid water from ice, is the influence of time scales on molecular structure. For solids, this influence is minimal. For liquids, it is so
important that the very definition of the term structure must incorporate it
in an essential way. I Consider a typical water molecule as it moves about

48

THE SURFACE CHEMISTRY OF SOILS

through the bulk liquid phase. On a time scale that is short compared with
a period of vibration for a hydrogen bond (about z x 10- 13 s), the water
molecule "sees" a spatial arrangement of its neighbors that is called the
instantaneous structure- (I structure). This I structure exhibits water molecules in a highly irregular arrangement because it exists on a time scale so
short that the position and orientation of individual molecules can be
momentarily far removed from their most probable values. Thus the
separation between a typical molecule and its nearest neighbors, as well as
the degree of hydrogen bonding among them, deviates considerably in the
I structure from the most stable average configuration. If the time scale
is lengthened so as to lie somewhere between 2 x 10- 13 s and the time
required for a water molecule to diffuse in the liquid through a distance
equal to its own diameter (about 10- 11 s) a typical molecule see a surrounding spatial arrangement called the vibrationally averaged structure
(V structure). This structure shows water molecules near their most
probable position because it exists on a time scale long enough to include
many hydrogen bond bending and stretching vibrations and therefore
represents an average over the positions of the molecules during those
vibrations. The V structure thus presents more local ordering and less
hydrogen bond distortion than the I structure.j Finally, on a time scale that
is very long compared with a diffusion time (about 10- 6 s), a typical water
molecule sees a spatial configuration of its neighbors known as the
diffusionally averaged structure (D structure). This structure includes all
effects of vibrational, rotational, and translational motion of the water
molecules. It is more ordered than the V structure because it represents
long-time averages of positions and orientations leading to only the most
probable molecular configurations. Since the time scale is long enough for
diffusive motions to take place in the liquid, a typical molecule does not see
just one set of neighboring molecules in the D structure, of course.
Instead, neighboring molecules come and go, leaving the' chosen typical
molecule to see the average or most probable sites and orientations they
occupy.
Even these brief remarks should make it clear that the concept of
structure in liquid water is a dynamic one. The molecular arrangements
perceived and particularly their degree of ordering very much depend on
the time scale involved. It is critically important to bear this feature in mind
when discussing the experimental methods used commonly to study the
structure of water since each method is itself characterized by a specific
time scale during which it probes a molecular environment. Several of
these experimental methods are indicated in Fig. 2.1. The molecular
structural parameters that can be deduced by applying the methods to
liquid water and aqueous solutions are listed in Table 2.1.
The infrared (IR) and Raman spectrometers available for studies of
liquid water cover a frequency range corresponding to periods of molecular
vibration between 10- 15 and 10- 12 S3. Thus optical spectroscopy can give
information concerning the transition from the I structure to the V

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

49

STRUCTURE
TIME SCALE IN
SECONDS (log)
I R SPECTRA
NEUTRON
SCATTERING
ESR SPECTRA
NMR SPECTRA
DIELECTRIC
RELAXATION
NEUTRON AND
X-RAY
DIFFRACTION
THERMODYNAMIC
PROPERTIES

Figure 2.1. Time scales for methods used to study the molecular structure of liquid
water.

Table 2.1. Molecular properties of liquid water and the experimental methods that
measure them

Method
Infrared and Raman
spectroscopy
Electron spin resonance

Incoherent neutron
scattering

Nuclear magnetic resonance

Dielectric relaxation

Neutron, electron, and


X-ray diffraction

Molecular parameter and


its physical significance
OR and hydrogen bond strength, orientation,
and length (V structure)
Solvation water molecule orientation Tc ,
correlation time for rotation of a solvation
complex
Ds , self-diffusion coefficient
TR ~ residence time for jump diffusion
TJ, correlation time for rotation of the dipole
moment
Water molecule orientation (NMR line shape)
TZ, correlation time for rotation about the
dipolar axis
T c , rotational correlation time for solvation
complexes
D s , self-diffusion coefficient
TJ, correlation time for rotation of the dipole
moment
a, measure of the spread of correlation times
about TI
Water-molecule 0 and H positions and bond
orientations (0 structure)

50

THE SURFACE CHEMISTRY OF SOILS

structure. The molecular parameters involved pertain to OR bonds in


water molecules and hydrogen bonds between water molecules, as indicated in Table 2.1.
Electron spin resonance (ESR) spectroscopy probes a molecular environment on the relatively narrow time scale of 10- 11 to 10- 10 sand
therefore gives information related to the transition between the V
structure and the D structure. The use of ESR is limited to aqueous
solutions containing solute atoms with unpaired electrons, i.e., molecular
spin systems that respond to an applied magnetic field." The induced
magnetization of these spin systems (e.g., Cu2+ or Mn2+), as well as its
subsequent relaxation upon removal of the perturbing magnetic field, are
affected strongly by the surrounding molecular environment. Thus one can
learn something about the structure of solvation complexes (e.g.,
CU(R20)~+) by observing the magnetic behavior of an appropriate set of
atomic electrons surrounded by liquid water molecules.
Incoherent neutron scattering (INS) can be used to study the translational, rotational, and vibrational motion of water protons on a time scale
between 10- 13 and 10- 10 S.5,6 Thus INS provides data pertinent to the V
structure and to the transition from the V structure to the D structure in
liquid water. The principal use of INS has been to characterize the
translational and rotational motion of water molecules through the interpretation of scattering data with model expressions. The three most
important model parameters used are the self-diffusion coefficient, D.,
which can also be measured in an experiment involving isotope-labeled
water molecules; 7 the residence time of a water molecule, 'TR, during which
it vibrates about a fixed position before "jumping" to its next position; and
the correlation time, 'Tl, which is a time constant for the decay of
correlation between the orientation of a water molecule at some initial
time and at some later time."
Nuclear magnetic resonance (NMR) spectroscopy probes a molecular
environment on the broad time scale between 10- 10 and 10- 3 sand
therefore provides data concerning the transition between the V structure
and the D structure plus data on the D structure itself. The physical basis
for the application of NMR to liquid water and aqueous solutions is
parallel to that for ESR, in that one studies the response of a magnetic
nucleus (e.g., a water proton) to an applied electromagnetic field." The
relaxation of the magnetization induced in a proton by a magnetic field is of
particular importance in liquid water studies. An analysis of relaxation
data with model expressions describing both intra- and intermolecular
proton-proton interactions leads to a calculation of the correlation time,
'T2, which may be interpreted as a decay time constant for correlation
between orientations of the proton-proton vector in a water molecule." For
solutes in aqueous solutions, it it also possible to determine the time
constant, 'Tc ' for the decay of orientation correlation of a solvation
complex through an analysis of NMR relaxation data.
Dielectric relaxation spectroscopy can probe the very broad time domain
between 10 12 lind 10 2 s. In principle. therefore, this method can be used

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

51

to obtain data on the V structure, the D structure, and the transition


between them. The physical basis of the method is connected closely to the
fact that the distribution of electic charge in a water molecule is primarily
dipolar. Thus, if a water molecule is subjected to an applied electric field,
the molecule will attempt to orient itself with its dipole moment pointing
along the direction of the field; if the field is time-varying, the molecule
also will attempt to follow changes in the field direction by reorienting itself
sympathetically. The extent to which this is possible, however, depends on
the constraints imposed on the molecule by its environment. Should the
molecule be coupled strongly with its neighbors, its ability to follow a
fluctuating electric field will be less than if the coupling were weak.
Whatever the degree of coupling, it is certain that the water molecule will
be less able to reorient in phase as the frequency with which an applied
electric field changes direction is increased, and at some frequency it will
fail to respond altogether. The mean frequency at which the failure
becomes complete is expected to be an indicator of the strength of the
bonds that constrain the water molecule. The greater this mean frequency,
the weaker the constraining bonds between the molecule and its environment. Thus dielectric relaxation spectroscopy consists of analyzing, with
model expressions, measurements of the complex dielectric permittivity as
a function of the frequency of an applied electric field. Typically, these
model expressions contain the decay time constant, T1, and a parameter a,
which equals zero if there is no spread of decay times about the mean value
9
T1 and tends to unity as the spread of decay times becomes larger.
Neutron or X-ray diffraction is the coherent, elastic scattering of either
neutrons or X-rays by atoms arranged ona periodic lattice. 6 ,10 Neutron
diffraction is analogous to the better known X-ray diffraction but differs
from the latter in two important respects. First, neutron diffraction
involves scattering by nuclei, whereas X-ray diffraction involves scattering
by atomic electrons. It follows that the scattering power of a given element
is different, in general, for the two processes. For example, deuterium,
having but one electron, has a low scattering power for X-rays but a high
(coherent) scattering power for neutrons. Second, neutron diffraction
probes a larger spatial domain in a target sample than X-ray diffraction
because of the greater penetration of neutrons into matter. Thus, structural information pertaining to relatively larger molecular units in sample
materials can be obtained. Since coherent scattering experiments require
the collection of data over long time periods, neutron and X-ray diffraction
give a picture of the structure of water on an effectively "infinite" time
scale. This picture, then, is an account of the D structure in liquid water
and aqueous solutions.
Thermodynamic data mentioned in Fig. 2.1 refer to the properties of
stable states and therefore give information pertaining to a time scale that
also is effectively infinite with respect to the dynamics of water molecules.
Besides this well-known fact, there are two important features of thermodynamic methods that set them apart from the other methods indicated in
Fig. 2.1. First. since the sole objective of chemical thermodynamics is the

52

THE SURFACE CHEMISTRY OF SOILS

development of exact mathematical relationships among the macroscopic


properties of a physical system, thermodynamic data cannot be interpreted
directly in terms ofmolecular structure. Thus the thermodynamic properties
of liquid water cannot provide unambiguous insights into the structural
behavior of water molecules. Second, the thermodynamic properties of
mixtures are only formally separable into properties that pertain to the
individual components. Any attempt to assign these properties to just the
water component in a solution or suspension must be understood as an
arbitrary action unless nonthermodynamic evidence exists to support it.
The molecular structure of liquid water is not yet a precise quantitative
concept despite the many studies carried out using the methods indicated
in Fig. 2.1. What has emerged from these studies over the past dozen years
is a firm qualitative-or perhaps semiquantitative-picture of the I, V,
and D structures that is reasonably self-consistent and sufficiently detailed
to serve as a basis for interpreting data on aqueous solutions and phyllosilicate suspensions.
The I structure in liquid water cannot be inferred from the experimental
methods listed in Table 2.1 because those methods provide data that are
time averages over many I structure configurations. However, the technique of molecular dynamics (MD) computer simulation has led to reliable
information about the I structure.lv-!' In this technique, a computer is
used to solve the classical mechanical equations of motion with a chosen
intermolecular potential function for a few hundred water molecules
constrained in space to maintaining the equilibrium liquid density, with
data on the instantaneous position and velocity of the molecules provided
both as numerical output and in the form of stereoscopic pictures. The
principal features of the I structure determined in this fashion are 12
1. A clear tendency for neighboring water molecules to orient themselves
into a tetrahedral, hydrogen-bonded structure
2. The absence of "clusters" that have mass densities very different from
the equilibrium liquid average
3. The absence of a monomer population and of structures resembling any
of the known ice structures
4. The existence of some non-hydrogen-bonded OH groups and of considerable distortion (nonlinearity) in OH 0 bonds
These characteristics are expected to be refined in the V structure of
liquid water in the sense that the positions of the water molecules become
more precisely localized and the distortions in the hydrogen bonds are
reduced. Both computer simulationvP and IR, Raman, and neutron
scattering experiments! confirm this expectation. In particular, there is a
significant narrowing in the distributions of nearest-neighbor distances and
OH . 0 bond angles, an increase in hydrogen bonding and in the number
of polygon structures with both even and odd numbers of molecules, and a
clear indication that the nearest neighbors of II typical molecule in the

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

53

b.

~.a. -0- ~
~~rf.P .P
-o..~~
-0".

(0)

( b)

Figure 2.2. (a) Ideal local tetrahedral configuration of molecules in liquid water.
(b) Monte Carlo computer simulation of the V structure in liquid water. (After
Rice 1)

liquid are arranged about it in a tetrahedral configuration, as illustrated in


Fig. 2.2.
The tetrahedral arrangement that characterizes the local ordering of
water molecules in the V structure dissipates after about 10- 12 s (at 283 K)
because the nearest neighbors of a given molecule are relatively free to
change their positions in response to thermal perturbations." This relaxation time is about one order of magnitude smaller than the time scale for
diffusive translational and rotational motion of the molecules in liquid
water at 283 K, as can be inferred from Table 2.2. The physical meaning of
Table 2.2. Correlation time constants for the diffusive
translational and rotational motions of a single molecule in liquid water"
T,K

273
278
283
288
293
298
303
308

D .. 10- 9 ms
2-1

1.092
1.313

1.543
1.777
2.021
2.302
2.620
2.919

'7"d,

psa

12.7
10.5
9.0
7.8
6.8
6.0
5.3
4.7

'7"1>

psb

17.9
14.9
12.6
10.8
9.3
8.1
7.2
6.4

'7"2,

psc

5.8
4.9
4.3
3.7
3.2
2.9
2.5
2.3

r" = 2a'/JD, (whcrc IJ ~ 0.144 nm is thc van der Waals radius of


H,O) is thc time required for a "diffusive step" in thc liquid,

Obtnincd from dielectric relaxation datu,


Ohlaincd from proton NMR relaxation data,
h

54

THE SURFACE CHEMISTRY OF SOilS

this fact is that the local environment of a water molecule in the liquid
undergoes many fluctuations during the course of an elementary "singleparticle" motion through translation or rotation, Therefore, the parameters in Table 2.2, although of molecular significance, describe the
transition from V structure to D structure only in the broadest terms and
cannot be used to examine either structure in great detail because those
details are washed out by many fluctuations. 9
The D structure of liquid water as revealed by X-ray, neutron, and
electron diffraction experiments'r'" comprises water molecules hydrogenbonded in an extensive network that exhibits local tetrahedral ordering.
The persistence of the tetrahedral structure (Fig. 2.2) is exemplified by the
fact that the coordination number of a water molecule in the liquid,
obtained by integration of the first peak in the radial distribution function
determined from X-ray diffraction data, is equal to 4.4 throughout the
temperature range from the melting point to the boiling point. However,
the diffraction data also indicate that water molecules outside the shell of
nearest neighbors may deviate considerably from the configurations projected on the basis of, e.g., an ice-like ordering. Therefore, a large number
of distorted or broken hydrogen bonds exists in the liquid, and it is these
bonds that determine the time-dependent properties of water.
The structure of liquid water as reviewed here is actually a kind of
sequence of structures whose degree of ordering and connectivity increases
with the time scale of molecular observation. Perhaps the most succinct
definition of liquid water structure that encompasses all of the known
qualitative characteristics related to time scales is that given recently by
F. H. Stillinger;' "Liquid water consists of a macroscopically connected,
random network of hydrogen bonds, with frequent strained and broken
bonds, that is continually undergoing topological reformation. [The] properties of water arise from the competition between relatively bulky ways
of connecting molecules into local patterns characterized by strong
bonds and nearly tetrahedral angles and more compact arrangements
characterized by more strain and bond breakage."
2.2. ELECTROLYTE SOLUTIONS

The influence of an ion in an aqueous electrolyte solution on the structure


of liquid water can be pictured spatially as a localized perturbation of the
tetrahedral configuration shown in Fig, 2.2. 15 In the ~()n of t~Lq\lig
n~arej:JM.jQQ.2.Jhe water molecules ar~ dominateci. biaae'5~..j~le.ctr{)
siric,!~~L!J,l!;l.~.l::~l!~:crt1ie-pri"!iji.~~l~ation sheil."tii'the~nexro~'ter region, the
water molecules interact weakly WIth the ion and form a structure known
as the secondary solvation shell or second-zone structure. Beyond the
second zone, the structure of the liquid is indistinguishable from that in the
pure bulk phase.
The study of aqueous electrolyte solutions with the methods listed in
Table 2,1 has as its objective the development of it molecular model of the

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

55

primary and secondary solvation shells. This model should include the
numbers and configurations of the water molecules in the shells, as well as
their residence times and the other single-particle parameters in Table 2.2.
All of these properties are expected to vary with the nature of the
electrolyte ions (particularly their valence and radius), with the concentration and temperature of the solution, and with the pressure applied to it.
The primary solvation shell of a small monovalent
cation appears to contain about six water molecules if the solution is dilute
and about three if the solution is concentrated. The residence time, Tb' of
the water molecules in the primary shell has been estimated on the basis of
dielectric and NMR relaxation experiments'? to be about 10 ps. This figure
is an order of magnitude larger than the residence time of a water molecule
in tetrahedral coordination with a given set of nearest neighbors in the bulk
liquid but is equal to the time required for a "diffusive step" (Table 2.2).
Thus the primary solvation shell of a small monovalent cation is relatively
well defined temporally, although diffusive exchange of solvation water
molecules with those in the bulk liquid is expected to be rapid. For larger
monovalent cations and for monovalent anions, relaxation experiments
indicate Tb < 5 ps, which is comparable with the residence time in the bulk
liquid. The uniqueness of the primary solvation shell for these ions is
therefore doubtful.
The number and orientation of the water molecules in the primary
solvation shell of Li+ have been investigated by X-ray and neutron diffraction. 15 ,17 In solutions of LiCI, the solvation shell contains about three
water molecules in a 10 m solution and about six in a 3.6 m solution. Over
the same range of molality, the angle (J between the dipolar axis of a
solvating water molecule and a coulomb field line passing through the
center of the molecule varies from 52 15 to 40 10. These angles are
near to but less than the 55 angle expected if lone-pair electrons in the
water oxygen atom interacted with the cation. If the solvating water
molecule behaved as a point dipole and aligned itself perfectly with the
cationic coulomb field, the angle observed would be 0 instead of 55.
Evidently the tendency toward dipolar orientation increases as solution
concentration decreases. The rotational correlation time, T2' for the solvating water molecules around u: in the limit of infinite dilution has been
estimated to be about 5 ps on the basis of the NMR relaxation method.l''
This is roughly the same as T2 for bulk water. For the rotational correlation
time of the entire solvation complex, T e , a value of 15 3 ps has been
determined by the same method.
The secondary solvation shell about an ion can be studied by neutron
diffraction and incoherent neutron scattering. 15 ,17 ,18 When applied to 5 m
LiCI, these methods, indicate that Li+ does not have a secondary solvation
shell. The same result would be expected for larger inorganic monovalent
cations. The absence of a secondary solvation shell around monovalent
ions is not surprising given the relatively small values of Th and T., quoted
MONOVALENT IONS.

THE SURFACE CHEMISTRY OF SOILS

56

above. On the other hand, neutron scattering results for dilute solutions
are not yet available, and it is possible that INS data will indicate the
existence of a weak second-zone structure around widely spaced Li +
cations, in agreement with the predictions of Monte Carlo computer
simulations of Li+-water systems.l"
On the basis of NMR relaxation experiments.l" the
residence time of a water molecule in the primary solvation shell of a
bivalent cation has been estimated to lie in the range 10- 9 s < Tb < 10- 4 s.
This time scale is much longer than that pertaining to self-diffusion in the
bulk liquid, and it implies that the solvation molecules move through a
solution of bivalent cations right along with the cations. For these water
molecules, both T2 and T c fall in the range 10 to 30 ps. Thus the primary
solvation shell is a temporally well-defined structural unit in aqueous
solutions containing bivalent cations.
The number of water molecules in the primary solvation shell of a
bivalent cation as determined by diffraction methods is always between six
and eight unless either cation size or ion-pair complexes intervene to
produce smaller values. 15 ,17 Thus the primary shell can be either an
octahedral or cubic complex. Table 2.3 shows how the solvation number
and orientation of water molecules in a solvation complex can vary with
electrolyte concentration. 17 The variation in (J with the molality of NiH is
striking. Evidently lone-pair interactions between a water molecule and
this cation are favored in concentrated solutions but dipolar interactions
are favored in dilute solutions.
Diffraction experiments also give evidence for a secondary solvation
shell around bivalent cations.Pr" This shell contains about 15 water
molecules whose mobility varies with electrolyte concentration. INS data 18
indicate clearly that the self-diffusion coefficient of the water molecules in
the second-zone structure approaches the diffusion coefficient of the
solvated cation as the molality of the solution decreases. Thus the
secondary solvation shell moves as a solvation complex with the cation in
dilute solutions.
BIVALENT CATIONS.

Table 2.3. Concentration effects on the number and


orientation of solvation water molecules in NiClz
solutions'?

Molality
, 4.41
3.05
1.46

0.85
0.46
11.1186

Solvation number
5.8
5.8
5.8
6.6
6.8
6.8

0.2
0.2
0.3
0.5
II.S
II.S

Orientation angle, (}
42
42
42
27
17

8
8
8
10
10
on 20

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

57

This brief review of the structure of water in electrolyte solutions is


intended to emphasize three points that will be important in understanding
the behavior of water molecules near clay mineral surfaces:
1. The effect of a cation on the structure of bulk water is localized to a
suite of no more than 6 to 20 solvation water molecules, even for
bivalent cations in a dilute solution. This local character of the
cation-water interaction, which has been observed clearly in MD
computer simulations.r" indicates that the primary solvation shell
screens the coulomb field of the cation very efficiently.
2. The primary solvation shell of a monovalent cation contains between
three and six water molecules that exchange relatively rapidly with the
surrounding bulk liquid. A secondary solvation shell, if it exists, is very
weakly developed.
3. The primary solvation shell of a bivalent cation contains between six
and eight water molecules that move with the cation as a unit. A
secondary solvation shell containing about 15 water molecules develops
as the cation concentration decreases and moves with the cation as a
unit.

2.3. WATER NEAR PHYLLOSILICATE SURFACES


Phyllosilicates bear siloxane surfaces on their basal planes and hydroxyl
groups along with Lewis acid sites on their edge surfaces, as described in
Sec. 1.2. The consensus at present is that the basal plane surfaces have the
greater potential to alter the configuration of water molecules from what
exists in the bulk liquid phase, and it is on siloxane surfaces that most
experimental studies of clay-water systems have focused. There are two
distinguishing properties of siloxane surfaces that are expected to affect the
molecular structure of a contiguous liquid water phase: (1) the nature of
the charge distribution on the siloxane ditrigonal cavities and (2) the nature
of the complexes formed between these surface functional groups and
cations. If there is no isomorphic cation substitution in a phyllosilicate
layer, its siloxane ditrigonal cavities expose only lone-pair electrons on
their oxygen atoms and their functional group reactivity has a very soft
Lewis base character. Surface complexes with cations are not stable in this
case. If there is isomorphic substitution in a phyllosilicate, the charge
distribution on its siloxane ditrigonal cavities depends on the location of
this substitution (Sec. 1.2). Isomorphic substitution in the tetrahedral sheet
localizes charge on the ditrigonal cavities more than does substitution in
the octahedral sheet. All of these features-the presence or absence of
unbalanced negative charge and the degree of charge localization-should
playa role in the organization of water molecules near siloxane surfaces.
The same is true of the valence and size of the cations complexed by the
siloxanc ditrigonal cavities bearing unbalanced negative charge. The
cationic coulomb field is determined by these two properties, as well as by

THE SURFACE CHEMISTRY OF SOilS

58

the inner- or outer-sphere character of the surface complex. Indeed, the


existence of an outer-sphere surface complex implies that a solvation shell
has formed around the complexed cation, and this is itself a perturbation of
the structure in bulk liquid water.
In this section, how surface charge distribution and surface cation
complexation in siloxane ditrigonal cavities affect the structure of water
is illustrated. The kaolinite group minerals are used to show how an
electrically neutral siloxane surface affects the molecular arrangement in
bulk liquid water. The vermiculite and smectite group mirierals are then
considered to clarify the effect of the isomorphic substitution pattern on
the configuration of water molecules whose "unperturbed" structure is
that in an aqueous electroyte solution, since these two clay mineral groups
form surface complexes with cations. Besides these inferences concerning
molecular structure, an estimate of the spatial extent of the influence of the
siloxane surface on the properties of water is considered. The spatial
domain wherein the water molecules take on a D structure that differs
measurably from that in bulk water or in an aqueous electrolyte solution of
the appropriate molality defines the region of adsorbed water on a siloxane
surface. The solvent properties of adsorbed water are expected to be
different from those of bulk-phase water and therefore to affect surface
functional group reactivity differently.
As mentioned in Sec. 1.1, kaolinite group
minerals exhibit insignificant layer charge in most cases, the upper limit
being x = 0.012 (about 0.02 mole kg-I). In many natural kaolinites, 2: 1
phyllosilicates are present in small amounts that nevertheless contribute
importantly to the apparent layer charge of the samples (Table 1.3). This
kind of contamination makes the unambiguous interpretation of surface
chemistry experiments very difficult, if not impossible. Perhaps a good rule
of thumb'" is to regard as suspect any sample of a kaolinite group mineral
whose cation exchange capacity at pH 7 is larger than 0.02 mole kg-I.
The D structure of water on the basal plane surfaces of kaolinite group
minerals is usually studied using halloysite, which in its dehydrated
form (7-A halloysite) is a disordered polymorph of kaolinite.F The
hydrated form (10-A halloysite) has the structural formula
[Si4 ] (AI4)OlO(OH)s . 4H zO and exhibits a basal plane spacing of 1.01 nm.
The four structural water molecules per unit cell can be removed either by
placing the sample in an atmosphere of relative humidity below 30 per cent
or by heating it at 60 to 70C. This removal of the water molecules is quite
irreversible kinetically and rehydration does not occur even after exposure
to 100 per cent relative humidity for two months. The basal plane spacing
of 7-A halloysite is precisely 0.72 nm, which means that the interlayer
separation in lO-A halloysite is 0.29 nm-exactly equal to the van der
Waals diameter of the water molecule.'} Accordingly, lo-A halloysite can
be envisioned as a disordered kaolinite with a monolayer of adsorbed
water in the interlayer space. On the basis of X-ray diffraction data. it has
KAOLINITE GROUP MINERALS.

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

59

been suggested that this monolayer comprises water molecules in the


hexagonal network configuration shown in Fig. 2.3.12 ,23 The stability of the
network derives from hydrogen bonding, both between the molecules and
between them and the atoms in the basal planes that enclose the interlayer
region. Because of the constraints imposed by the requirement of four
water molecules per unit cell and by epitaxy with the basal plane surface
atoms, the hydrogen bonds in the monolayer must be about 0.3 nm long
and those between the water molecules and the mineral surfaces must
deviate considerably from linearity. These structural characteristics are
similar to those of bulk liquid water, even to the extent of preserving local
tetrahedral ordering. That the hydrogen-bond network is more like that in
liquid water than that in ice with respect to structural ordering also has
been concluded from IR spectroscopic studies.e"
Given the structure in Fig. 2.3 as a prototype of the monolayer
arrangement of adsorbed water on kaolinite group minerals, there remains
only the question of its "single-particle" properties. Proton and deuteron
NMR measurements are instructive in this respect." As indicated in
Table 2.1, NMR line shapes show whether a preferred orientation of water
molecules exists on the NMR time scale, and NMR relaxation data can
yield a value for the rotational correlation time, 72' In the case of lO-A
halloysite, the NMR signal from water protons or deuterons in the

Figure 2.3. The D structure of water in the interlayers of


Hendricks and Jefferson 23 )

1
I
I

- -M---'--'t)-e----

~ ~------

-\ -----1--\
\

I
I

lO-A

halloysite. (After

60

THE SURFACE CHEMISTRY OF SOilS

adsorbed m<:?nolayer (as opposed to those in "micropore" water or in clay


mineral OH groups) consists of a single peak, just as is found for bulk
liquid water ,1 indicating no preferred orientation relative to the basal
planes on th~ NMR time scale. This result means that the water molecules
rotate quick'y enough to sample all possible orientations in the hexagonal
network dur~ing the time frame in which NMR probes their structure. An
analysis of:rVMR relaxation data for water protons and deuterons on 10-A
halloysite24t-eads to the conclusion that TZ = 0.144 exp(2315/T) , where TZ
is in picosecc;:mds and T is absolute temperature. At 298 K, TZ = 340 ps, or
about two o:::rders of magnitude larger than TZ in liquid water at the same
temperature (Table 2.2). The same kind of discrepancy is found for the
correlation J:ime, Tl' if the NMR relaxation data are interpreted with a
model descr_bing a water molecule rotating rapidly around its dipolar axis
while that a~is itself tumbles about an axis parallel to the basal planes. Z4
At 298 K, 1"':1 = 0.2 /LS, according to this estimate, a value four orders of
magnitude larger than Tl in bulk water at the same temperature. A
confirmatiors of the NMR value for Tl has been established from direct
measuremersts of this correlation time for water on kaolinite by dielectric
relaxation sI?ectroscopy. Z5 At monolayer coverage and 298 K, Tl = 4 ILs.
The static di electric constant of the water is about 2.0, far below the value
for liquid w.,.ter but near that of ice Ih(ordinary hexagonal ice). Similarly,
Tl = 2 ILS for ice Ih at 250 K, and a value near 3 ILS is found by
extrapolating the temperature dependence of Tl for ice Ih to 298 K.
However, tlJ.e decay-time spread parameter, a (Table 2.1), is less than 0.03
for either liquid water or ice Ih, whereas for water on kaolinite,
0.5 < a < (1.6, indicating a broad spread of decay times and therefore a
broad range' of local molecular environments.F
The struvtural concept that emerges for the water monolayer on
kaolinite is that of a strained, hydrogen-bonded network whose D structure reflects a compromise between the topological disorder of liquid water
and the rigid configuration demanded by strict epitaxy with the atoms in
the basal pl~nes of the clay mineral. The monolayer is liquid-like in not
exhibiting a preferred orientation of its constituent molecules on the NMR
time scale and in comprising relatively weak, distorted hydrogen bonds.
However, tlJe clay mineral surface is able to slow down the single-particle
motions in the monolayer considerably, as evidenced by the fact that the
rotational correlation times, Tl and TZ, have values appropriate to solid
water. These structural characteristics are underscored by the temperature
dependence of the apparent specific heat capacity of the monolayer
water. 24 ,Z6 1his quantity is defined by the equation
(2.1)
where Cp is the heat capacity at constant pressure of a mixture containing
m.; kilograms of water and mk kilograms of kaolinite, c'k is the specific heat
capacity of fully dehydrated kaolinite, and ~c is the apparent specific heat
capacity of t"'c water. As mentioned in Sec. 2.1, this method of partitioning

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

61

a thermodynamic property of a mixture into terms pertammg to the


components has only a formal significance and caution must be exercised
when structural interpretations are made of functions like 4>c in {he
absence of supporting nonthermodynamic data. In the case of the wateron
lO-A halloysite, the fact that 4>c closely follows the temperature dependence of the specific heat capacity of ice Ih for T s 150 K, rises quickly to a
sharp maximum at 260 K, and then falls to equal the specific heat capacity
of liquid water at 273 K is consistent with a water monolayer containing
molecules whose rotational and translational motions are hindered by
interactions with the clay mineral surface but whose structural disorder
resembles that in liquid water even though it is actually induced partly by
the demands of epitaxy.
The effect of the kaolinite surface on the structure of water beyond
monolayer coverage has not been ascertained conclusively. Multilayers of
water are known to form on the clay mineral at relative humidities greater
than 20 per cent, and about 10 molecular layers are thought to exist at
relative humidities near 98 per cent. 27 Thus an absolute upper limit for the
dimension of the region of adsorbed water in a kaolinite suspension should
be around 3 nm. Data on the thermodynamic properties (heat of immersion, partial specific entropy, isosteric heat of adsorption) of Li-saturated
kaolinite-water systems, which do not show hysteresis, indicate consistently that differences between bulk liquid water and water on the clay
are detectable up to about three molecular layers of coverage.r" Perhaps
the lower limit for the dimension of the adsorbed water region is then
about 1 nm.
The D structure of water molecules on
the siloxane surfaces of vermiculite group minerals has been investigated
only for trioctahedral vermiculites with the general structural formula'?
VERMICULITE GROUP MINERALS.

M x[SiaA I8 - a](MgbFe(III)c Fe (II)c'TidAI6-b-c-c' -d) 020( OH)4

where M represents one mole of cation charge on the basal planes and
x = 2

+ b + c' - a - d

(2.2)

is the layer charge, defined conceptually in Sec. 1.1 and given numerically
in Table 1.3. The trioctahedral vermiculites offer the distinct advantage of
being available in macrocrystalline forms of high purity that permit
exhaustive study by X-ray diffraction. The interlayer water structures
determined for these clay minerals are presumed to be models for those
that exist on the surfaces of the more disordered dioctahedral vermiculites
found commonly in soils.
The configuration of water molecules between the basal planes in
vermiculite depends sensitively on the nature of the 'cation complexed on
these surfaces to balance the negative charge produced by isomorphic
substitution of Al,l for Si4 + in the tetrahedral sheet. Because of this
substitution. the siloxanc ditrigonal cavities exhibit a relatively localized
j

62

THE SURFACE CHEMISTRY OF SOILS

charge distribution that can interact strongly with both cations and water
molecules, as mentioned in Sec. 1.2. The overall effect of the charge
localization should be a proximity of the complexed cations to tetrahedral
sites containing At3+ and the formation of relatively strong hydrogen
bonds between interlayer water molecules and surface oxygen atoms.
Aside from these effects of the vermiculite surface, however, it is expected
that the organization of water molecules in the interlayer region conforms
to the behavior observed in relatively concentrated aqueous solutions,
described in Sec. 2.2. In particular, water molecules in a fully hydrated
vermiculite should be coordinated in a single solvation shell about monovalent cations and in two shells about bivalent cations unless stereochemical
factors intervene to make inner-sphere surface complexes with the cations
more stable than outer-sphere complexes. Moreover, the solvation water
molecules near monovalent cations should be relatively mobile and those
in the first solvation shells about bivalent cations should move with the
cation as a unit. Reasoning on the basis of the known behavior of water
molecules in aqueous electrolyte solutions, one predicts that hydrated
vermiculites comprise interlayer ionic solutions, with the cations and
solvating water molecules influenced by the coulomb fields emanating from
tetrahedral sites containing AI3+.
In atmospheres of relative humidity greater than 20 per cent, Mgvermiculite forms a hydrate with a basal plane spacing of 1.436 nm, which
is adequate for the accommodation of two monolayers of water molecules
in the interlayer region.i" The arrangement of the Mg cations and the
water molecules has been studied intensively by X-ray diffraction, and the
details of the cation and water-molecule oxygen atom positions have been
established.l" An illustration of the structure of the interlayer region is
given in Fig. 2.4. The magnesium cations are positioned midway between
opposing basal planes over sites in the tetrahedral sheet that contain A13+ .
This arrangement permits a localized charge balance between the clay

Figure 2.4. The D-structure of water in the interlayers of the two-layer hydrate of
Mg-vermiculite. (After Alcover and Gatineau30 )

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

63

mineral surface and the Mg cations. Each Mg2 + is surrounded by f.wo


solvation shells of six water molecules each to give a total solvation number
of 12. The molecules in the primary shell are located 0.206 nm from the Mg
cation and coordinate to the latter through lone-pair electrons. The
molecules in the secondary shell are about 0.49 nm away from the cation.
All of the solvating water molecules can form hydrogen bonds, both
between shells and with the basal plane oxygen atoms. These bonds are
about 0.29 nm in length. This configuration of the water molecules is
strikingly like that observed around bivalent cations in concentrated
aqueous solutions. In particular, NiH (r = 0.069 nm), which has about the
same ionic potential as MgH (r = 0.072 nm), also coordinates six water
molecules through lone-pair electrons (Table 2.3) in a primary solvation
shell at 0.207 nm. 31 It appears from this comparison that the two-layer
hydrate on Mg-vermiculite is indeed similar to what is found in a
concentrated aqueous solution.
Magnesium-vermiculite also forms monolayer hydrates with basal plane
spacings of 1.163 and 1.153 nm. 29 ,30 These hydrates are distinguished by
the configuration of the Mg2 + solvation complexes (outer-sphere surface
complexes) in them. The hydrate with the larger basal plane spacing
contains MgH in the centers of flattened tetrahedra formed by water
molecules; the other hydrate contains MgH at the apex of a pyramid
whose base comprises three water molecules.
Other bivalent cations tend to form solvation shells on the vermiculite
surface in a manner similar to MgH, but structural differences can arise
because of differing Lewis acid character (based on ionic potential and
polarizability) and stereochemistry. 30 For example, the single monolayer
hydrate of Ba-vermiculite contains BaH partially inside the siloxane
ditrigonal cavity, with six water molecules placed in the midplane of the
interlayer region in one-to-one correspondence with the oxygen atoms of
the ditrigonal cavity (i.e., an inner-sphere surface complex). This arrangement evidently reflects the close fit of BaH (r = 0.136 nm) in the
ditrigonal cavity (r = 0.13 nm) and the softer Lewis acid character of the
cation, which gives it less affinity for the very hard Lewis base H 20. As
another example, the two-layer hydrate of Ca-vermiculite can contain both
octahedral and cubic coordination of the water molecules solvating the
Ca 2 + . This possibility, which also exists in aqueous solutions;" evidently is
the result of a combination of a relatively high ionic potential and a large
radius ratio with 0 ( - II).
Considerably less information is available concerning the D structure of
water molecules hydrating vermiculites that form surface complexes with
monovalent cations. It appears that both Li- and Na-vermiculite can form
mono- and bilayer hydrates, whereas K-, Rb-, and Cs-vermiculite cannot.
For the latter, inner-sphere surface complexes are stable against solvation
of the cations because of Lewis acid character and stereochemistry. These
factors are especially notable when comparing K + with Ba2+, which have
almost identical ionic radii. Barium-vermiculite contains a monolayer of

THE SURFACE CHEMISTRY OF SOILS

water molecules even when its Ba2+ ions are in inner-sphere complexes,
evidently because of its higher ionic potential.
The two-layer hydrate of Na-vermiculite contains Na+ with octahedral
primary solvation shells.F' Because the cation is monovalent, all possible
surface complex sites are filled and no more than six water molecules can
solvate a cation uniquely. (By comparison with data on the solvation of
Na+ in aqueous solutions, one would conjecture that only a primary
solvation shell exists even if all surface complexation sites are not filled.)
Thus the interlayer region of the two-layer hydrate consists of a network of
solvation octahedra with hydrogen bonding both within the octahedra and
with surface oxygen atoms of the clay mineral. 33 These hydrogen bonds are
somewhat longer than those in the two-layer hydrate of Mg-vermiculite
and therefore are expected to be weaker. At relative humidities below
40 per cent, Na-vermiculite forms a monolayer hydrate wherein Na"
coordinates to surface oxygen atoms in one basal plane and rests on three
water molecules bonded to the opposing basal plane in a fashion similar to
what occurs in the 1. 153-nm monolayer hydrate of Mg-vermiculite. Model
calculations of the arrangement of water molecules in a hypothetical
monolayer hydrate of K-vermiculite suggest that the interlayer water
structure is like that in the monolayer hydrate of Ba-vermiculite, described
above.i" Thus, to some extent, parallels can be drawn between the structures of the solvating water molecules in monovalent and bivalent cationsaturated vermiculites.
The dynamic properties of the interlayer water structures in Ca-, Na-,
and Li-vermiculite have been investigated by INS. 35 The data for the twolayer hydrate of Ca-vermiculite were modeled quantitatively under the
assumptions (a) that a fraction of the adsorbed water protons are in the
octahedral primary solvation shells of the Ca2+ cations and are immobile
on the neutron scattering time scale and (b) that some of the remaining
adsorbed water protons diffuse by jumps within a region bounded by the
opposing siloxane surface and the solvated Ca2+ while others undergo
isotropic, translational jump diffusion between adjacent bounded regions.
The two jump-diffusion residence times were found to be about 15 ps and
150 ps. Assumption (a) is in agreement with the residence time of 10- 9 to
10- 4 s observed for primary solvation shell water molecules in aqueous
solutions containing bivalent cations. Moreover, since the diffusion coefficients of bivalent cations on vermiculite." are around 10- 12 m2s-1, they
can move a distance equal to their own diameters (= 0.2 nm) in about
10- 8 s, which is also much longer than the neutron scattering time scale
(Fig. 2.1). Therefore, water protons in the primary solvation shell should
be immobile targets for neutrons. Assumption (b) is consistent geometrically with the D structure in Fig. 2.4. The effective self-diffusion coefficient
for the water protons in the region between solvation shells was determined to be about 10- 9 m2s- 1 , about one half the value found in bulk
liquid water (Table 2.2).

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

65

Neutron scattering data for Li- and Na-vermiculite, on the other hand,
gave no indication of water protons being immobile on the neutron
scattering time scale. 35 This result is consistent with the behavior of water
molecules in aqueous solution, since the residence time in the primary
solvation shell of a monovalent cation is about 10- 11 s, well within the time
scale probed by neutrons. However, as shown in Table 2.4, the selfdiffusion coefficients of water molecules on Li- and Na-vermiculite were
found to be much smaller than the bulk liquid value at 298 K. These data
suggest that, even in the two-layer hydrate, the solvating water molecules
exhibit only about 5 per cent of the mobility they have in the bulk liquid
phase and about 10 per cent of that in the primary solvation shell of a
monovalent cation in aqueous solution (D s = 1.3 x 10- 9 mZs- 1) 16 . This
reduction in water molecule mobility is evidently produced by interactions
with the charge distribution on the siloxane surface.
Magnetic resonance (ESR and NMR) studies''? have provided additional
details about the orientational motion of the water molecules adsorbed by
vermiculites. ESR spectra of Cu-vermiculite and NMR spectra of Mg- and
Na-vermiculite indicate clearly that the primary solvation shells of the
cations on the two-layer hydrate are octahedral complexes with a preferred
orientation relative to the siloxane surface. For Cu-vermiculite, the
symmetry axis through the solvation complex, Cu(HzO)~+, makes an angle
of about 45 with the siloxane surface; on Na-vermiculite the axis through
Na(HzO)t makes an angle of 65. The value of T e , the correlation time for
the rotation of Na(HzO)t around its symmetry axis, is about 10- 7 s at 298
K. This value is four orders of magnitude larger than T e for a solvation
complex around a monovalent cation in aqueous solution. 16 Not quite as
disparate are TZ for Na(HzO)t, equal to 100 ps at 298 K, and TZ for a
monovalent solvation complex in dilute aqueous solution, equal to about
5 ps at the same temperature. These data show that the siloxane surface
retards the orientational motion of the water molecules.
The spatial extent of the adsorbed water layer on vermiculite group
minerals has been estimated on the basis of thermodynamic properties and
self-diffusion coefficients for water on Li- and Na-vermiculite.P" The
Table 2.4. Self-diffusion coefficients for water on vermiculite
determined by INS 35

Exchangeable
cation
Li
Li
Li
Na
Na
Na

Water content,
kg H 20/kg clay

d(OOl),
nm

0.079
0.114
0.135
0.080
0.111
0:185

1.275
1.33
1.405
1.371
1.392
1.448

0.1
0.5
1.1

0.6
0.6
1.7

66

THE SURFACE CHEMISTRY OF SOILS

self-diffusion coefficients have values essentially equal to the bulk water


value for basal plane spacings larger than 5 nm, which therefore can be
taken as an upper limit for the dimension of the adsorbed water region.
The structure of water adsorbed by smectite
group minerals has been studied extensively in both its static (D structure) and dynamic aspects.i'? As with water molecules on vermiculite, the
behavior of water on smectite surfaces is conditioned sensitively on the
type of exchangeable cation and on the location of isomorphic cation
substitutions in the layer structure. In many respects, a discussion of the
configuration of water molecules hydrating smectites is parallel to that for
vermiculite.
Almost all of the experimental data pertaining to the structure of the
one-layer hydrate of monovalent-cation-saturated smectites can be interpreted in terms of the molecular arrangement in montmorillonite illustrated in Fig. 2.5. 40 This is a strained, ice-like configuration of water
molecules around the exchangeable M+ cations, with bonds formed both
intermolecularly and with the ditrigonal cavities in the smectite surface.
The nearest-neighbor distance between water molecules is 0.32 nm, and
five molecules are assigned to each exchangeable cation in the completed
monolayer. Since some of the water molecules have a hydroxyl group
proton inside a ditrigonal cavity on the siloxane surface, the network of
hydrogen bonds in the water molecules is broken in places. These defects
permit a variety of reorientational motions that are not possible in ice Ih.
The principal lines of experimental evidence that support the water
structure portrayed in Fig. 2.5 are as follows.
SMECTITE GROUP MINERALS.

1. Studies of X-ray diffraction by the water oxygens and of neutron


diffraction by the water protons have shown that the molecules in the
one-layer hydrate are positioned along the crystallographic c axis
0.55 nm from the aluminum ions in the octahedral sheet of montmorillonite;" in agreement with the structure in Fig. 2.5. Moreover,
the lateral position of the water molecules is correlated strongly with the
oxygen atom arrangement on the siloxane surface, with some molecules
entrained over the ditrigonal cavities. However, the scattering density
profile for the hydrate does not exhibit features consistent with a highly
ordered D structure. Measurements of the apparent specific heat
capacity (Eq. 2.1) support this conclusion, in that they show cPc
following the temperature dependence of the specific heat capacity of
ice Ih for 100 K < T < 150 K, then rising toward the value for the
specific heat capacity of liquid water at 273 K. 42 This behavior suggests
that the motion of the water molecules in the one-layer hydrate is not
highly restricted near 298 K.
2. Infrared spectroscopic studies'" give evidence for a three-molecule
solvation shell around the monovalent cations when the water content is
low. The water molecules are coordinated to the cations through

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

67

(a)

Figure 2.5. The D structure of water in the interlayers of the one-layer hydrate of
M+-montmorillonite (M = Li, Na, K, etc.). Basal plane oxygens are shown as
shaded circles. (a) View along an axis perpendicular to the crystallographic c axis.
(b) View along the c axis, with water molecules nearest the upper basal plane (not
shown) indicated by dashed lines. (After M amy 4fJ)

lone-pair orbitals, and one of the water protons is directed along the
crystallographic c axis into a siloxane ditrigonal cavity. If a significant
localization of surface charge exists because of isomorphic cation
substitution in the tetrahedral sheet, however, hydrogen bonds are
formed between the solvating water molecules and surface oxygen
atoms, as in the vermiculite group minerals.
3. INS 44 and dielectric relaxatiorr'" studies both indicate that the water
molecules solvating the monovalent exchangeable cations on montmorillonite are roughly as mobile, in respect to translational and
reorientational motion, as water molecules in the bulk liquid. For
example, the INS data show that no water molecule is stationary On the
neutron scattering time scale. The data can be described mathematically
by a model that includes both jump translational and rotational
diffusion. In the one-layer hydrate of Li-montmorillonite, D, =
4 X 10- 10 m 2s-I, with a jump distance of about 0.35 nm, and 'Tl =
15 ps at 293 K. The values of D; and 'Tl suggest a sluggish motion of
the water molecules, consistent with what is observed for solvating
water molecules around monovalent cations in aqueous solutions.
The values of 7"1 for Na- and K-montmorillonites hydrated by a monolayer of watcr are around 5 ps, according to dielectric relaxation

68

THE SURFACE CHEMISTRY OF SOilS

measurements.i" with a broad spread of relaxation times about the


mean value (a = 0.7; Table 2.1.). The spectrum of relaxation times is
consistent with a defective network of hydrogen bonds like that shown
in Fig. 2.5. Proton and electron spin magnetic reasonance studies37 39
tend to support these data by indicating 72 values around 100 ps and
3
7 c values near 10 ps.
The structural characteristics of water on bivalent-cation-saturated
smectites have not been worked out in enough detail to permit the
development of a schematic drawing like that shown in Fig. 2.5. INS data
for the two-layer hydrates of Ca- and Mg-montmorillonite'v'" can be
interpreted with the model used for Ca-vermiculite: an assembly of
octahedral solvation shells with slowly diffusing water molecules
(D; = 3.4 x 10- 10 m 2s- 1) interspersed between them. The residence time
and jump distance for the diffusion of water molecules between the
intersolvation shell regions are 100 ps and 0.5 nm, respectively. The jump
distance agrees well with the value of the mean distance between solvation shells calculated on the basis of specific surface area and surface
charge density of the clay mineral.l Dielectric relaxation spectroscopy
investigations'f indicate that 71 = 1 /LS, which is about the same as in ice
Ih but five orders of magnitude larger than in bulk liquid water. All of
these values suggest that the interlayer region is organized more or less as
depicted in Fig. 2.4, i.e., like a two-dimensional aqueous solution with
hydrogen affected by the charge distribution on the siloxane surface.
The spatial extent of adsorbed water on smectite surfaces is a matter of
some controversy. Infrared spectroscopy, NMR relaxation, and X-ray and
neutron diffraction experiments all point to a thickness of the adsorbed
water film of around 1.0 nm.39 However, certain thermodynamic data,
summarized for Na-montmorillonite in Table 2.5, suggest a thickness as
great as 10 nm or more.t" These data are for partial and apparent specific
properties of montmorillonite-water systems whose variation with water

Table 2.5. Dependence of some thermodynamic properties of Na-montmorillonite-water mixtures on water content"
Property a
Partial specific volume
(iJcPV/dT)P.8
-( iJcPV/dPh.8
Apparent specific
heat capacity == cPc
(iJcPc/iJPh.8
(iJcPs/iJPh,8

(acPs / a(Jv h.8

Comparison with value


in bulk liquid water

Behavior of value with


increasing water content

Larger
Larger than (iJvw/ iJ T)p
Smaller than -(iJvw/iJPh

Decreases
Decreases
Increases

Larger
Larger than (iJew/ iJPh
Smaller than (iJsw/dPh
Larger than (iJsw/iJvwh

Decreases
Decreases
Increases
Decreases

"'v - apparen: Npeclflc volume: "'~ - uppurenl Npeciflc entropy,

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

69

content has been interpreted structurally in terms of the water component


alone. The dangers inherent in this practice were alluded to in Sec. 2.1, and
this point alone may be enough to settle the issue of disagreement between
direct structural methods and thermodynamic approaches. Indeed, one
alternative for removing the disagreement is to assume that partial and
apparent specific quantities reflect changes in the smectite (particularly the
exchangeable cations) as well as in the water and that the former changes
become dominant as water content increases. This point of view cannot be
rejected in the present absence of conclusive experimental evidence
against it. On the other hand, if partial and apparent specific properties
represent only the adsorbed water, what can be said to bring the results of
thermodynamic measurements into agreement with other data?
For spectroscopic data, a difference in time scales can be indicated.
Thermodynamic data refer to effectively "infinite" scales of time and
space, characteristic of the D structure, whereas spectroscopic measurements probe either the V structure or some predecessor of the D structure.
The degree of ordering is less for these latter structures because the time
scale on which they exist is shorter, and thermodynamic properties may
truly reflect a multiplicity of cooperative interactions that can be perceived
only after a long-time measurement. This still leaves a question concerning
neutron and X-ray diffraction data, which refer to the same time scale as
the thermodynamic properties but appear to be in conflict with them.
It is known that the D structure in bulk liquid water, as deduced from
diffraction patterns measured at 273 K, differs only a little from that in ice
Ih, whereas the thermodynamic properties of liquid water at 273 K differ
greatly from those of ice Ih. If this comparison applies to adsorbed water as
well, then perhaps partial specific properties are more sensitive to structural changes than are diffraction patterns. It is also well known! that the
mere presence of a space-filling macromolecule in liquid water causes the
structure to be strained while strengthening some of the hydrogen bonds.
In effect, some of the topological freedom in the liquid disappears and the
fluctuations between bulky and compact networks of water molecules shift
to favor the bulky configurations. This picture could be an accurate
description, on the level of molecular structure, of how the trends listed in
Table 2.5 come about.
2.4. THE SOLVENT PROPERTIES OF ADSORBED WATER

The definition of adsorbed water adopted in Sec. 2.3 requires an arrangement of water molecules that differs significantly from that in an appropriate reference aqueous phase. For water on the surfaces of kaolinite group
minerals the reference phase is bulk liquid water, whereas for water on
vermiculite and smectite surfaces the reference phase is an aqueous
solution because of the presence of exchangeable cations on the 2: 1 layer
silicates. On the basis of this definition, the consensus developed in
Sec. 2.3 is that the spatial extent of adsorbed water on a phyllosilicate

70

THE SURFACE CHEMISTRY OF SOILS

surface is, conservatively, whatever is included in the region bounded by a


plane about 1.0 nm from the basal plane of the clay mineral. This sharp
geometric demarcation is, of course, an oversimplification because the
influence of the phyllosilicate surface on the configuration of water
molecules must decrease in a continuous fashion with distance, but a
bounding plane at 1.0 nm is expected to include all but a few per cent of
the siloxane surface effects on water structure per se.
With the exception of Li-vermiculite and certain low-layer-charge vermiculites which also swell in water, a 1.0-nm thickness for the zone of
surface influence and available X-ray diffraction data"? imply that the
interlayer water in vermiculite group minerals is always adsorbed water.
The same is true for smectite group minerals saturated with bivalent
exchangeable cations. However, if a smectite carries monovalent exchangeable cations, particularly Li+ and Na +, it will swell in water and the
interlayer region will fall within the defined zone of adsorbed water only
when free swelling is inhibited by ionic strength or relative humidity
control.F In the case of kaolinite group minerals, any interlayer water is
also adsorbed water because only a single monolayer is involved. The
water on the external surfaces of kaolinite exists within the adsorbed zone
for relative humidities below about 80 per cent."? The general conclusion
that can be drawn on the basis of the 1.0-nm criterion is that, unless free
swelling occurs, the interlayer water on phyllosilicates is structurally different from bulk liquid water or water in aqueous solutions, regardless of the
overall amount of water in a phyllosilicate-water mixture. Chemical reactions that take place in the interlayer region accordingly can be expected to
show some influence from a perturbed water structure.
One of the properties of liquid water that relate to its solvent characteristics is the dielectric constant, D, equal to the ratio of the static permittivity
to the permittivity of vacuum. At temperatures near 300 K, D = 80 for
bulk liquid water. 9 Measurements of the dielectric constant for adsorbed
water on phyllosilicates, despite its importance in surface chemical phenomena, are neither abundant nor conclusive because of difficulties
involved with separating out ionic surface conductance effects. The data
available suggest that D = 20, with values reported from 2 to 50. 25 ,40 ,45
These depressed values of D imply that adsorbed water molecules on
phyllosilicates are intrinsically less free to reorient along an applied static
electric field than are the molecules in bulk liquid water. This restriction
evidently comes about because of preferential orientation in water molecules that are imposed by the strong coulomb field of exchangeable cations
and the exigencies of hydrogen bonding. The most significant chemical
effect of this loss of orientational polarization and consequent lowered
dielectric constant in adsorbed water should be the enhancement of
complex formation, both between dissolved species and between exchangeable cations and siloxane ditrigonal cavities. It is well known that,
with a reduction in aqueous dielectric permittivity. ion association is
favored and the development of a diffuse electrical double layer ncar a
colloidal particle is retarded. 414

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

71

Another important chemical property of adsorbed water on vermiculite


and smectite surfaces is its Brensted acidity. This property should refer
principally to the acidity of the solvated exchangeable cations, as described
by the reaction
M(HZo)m+
= MOH(HZo)(m-1)+
+ H+
n
n-1

(2.3)

The equilibrium constant of this reaction has been measured for a variety
of metal cations in aqueous solutions and is known to correlate positively
with both ionic potential and Lewis acid softness.t? As the ionic potential
increases, the intensity of the positive coulomb field of the cation increases
and repulsion of a solvating water proton becomes more likely. As the
Lewis acid softness'" increases, the covalency of the M-O bond in a
solvation complex increases and electron density is withdrawn from the
O-H bond, thereby promoting the loss of the proton.
The reaction in Eq. 2.3 has been investigated extensively on siloxane
surfaces by coupling it with a Brensted base protonation reaction, e.g.,
M(HZo)m+
+ NH3
n

MOH(HZo)(m-1)+
+ NH+4
n-1

(2.4)

Table 2.6 shows experimental data pertaining to this reaction on


montmorillonite. 51 The last column of the table shows values of the
conditional equilibrium constant,

CK = {MOH(HzO)~~11)+}{NHt} =
(M(H zO):+)(NH3 )

{NHt}z
(HzO)(NH3 )

(2.5)

where the braces refer to concentration in moles per kilogram of clay

Table 2.6. Data relating to the protonation of ammonia on Wyoming montmorillonite'l'


Ionic
potential,"
nm""

Misono
softness, 50
nm

Li+

13.5

0.053

Na+

9.8

0.111

K+

7.3

0.189

Ca2+

20.0

0.165

Mg2 +

27.8

0.096

Exchangeable
cation

{NHt},
mol, kg"

mof

0.20
0.98
0.20
0.98
0.20
0.98
0.20
0.98
0.20
0.98

0.23
0.17
0.16
0.10
0.10
0.11
0.80
0.16
1.01
0.74

1.32b
0.15
0.64
0.05
0.25
0.06
16.0
0.13
25.5
2.79

Calcuhued u.11I1l dulU from 1'1I"le I. I.


hAil vlllue. culeulntcd f"r I kll ",' cllly under the 1Iumpuon Ihlll (NII.l
NII.lIII

cK,

Water
activity

~ 0.2 III 110,7%

.,,11111011 of

72

THE SURFACE CHEMISTRY OF SOilS

and the parentheses indicate thermodynamic activity. In Eq. 2.5, the


stoichiometry of Eq. 2.4 has been noted, and in calculating "K it has been
assumed that the activities of adsorbed water and ammonia are equal to
their relative vapor pressures. The values of {NHt} and "K at constant
water activity in Table 2.6 illustrate two general rulesr'" (1) for cations ofthe same Lewis acid softness (e.g., Na+ and Mg2+), protonation increases
with ionic potential, and (2) for cations of the same ionic potential,
protonation decreases with Lewis acid softness.
The other important feature of the {NHt} and CK values is their
increase, for any exchangeable cation, with decreasing water activity. The
magnitude of this increase appears to depend on both ionic potential and
Lewis acid softness, becoming larger as both parameters increase. The
implication of this trend is that a solvated exchangeable cation becomes
more acidic as the amount of adsorbed water decreases. Evidently, as the
number of water molecules on the siloxane surface is reduced to the point
where only primary solvation shells can form, the burden of screening the
charge of the exchangeable cations becomes so great that proton dissociation from the solvating water molecules increases significantly. Experimental support for this hypothesis has come from both NMR and IR
spectroscopic studies and from conductivity measurernents.P'Y At low
water contents on smectite and vermiculite group minerals, the mean
lifetime of a hydrated proton, THP+, is about 10- 10 sat 300 K and the mean
interval between associations of a water molecule with a proton, TH,o, is
10- 4 to 10- 5 s. The degree of dissociation of the adsorbed water is, by
definition.i" equal to the ratio TH)0+/TH 20 = 10- 5 to 10- 6 , In bulk liquid
water, 53 TH ) 0+ = 10- 12 sand TH 20 = 5 X 10- 4 s, which leads to 2 x 10- 8
for the degree of dissociation. Thus it appears that the degree of dissociation in adsorbed water is about two orders of magnitude larger than in bulk
liquid water. Model calculations of the distribution of positive charge on
water molecules solvating an exchangeable cation on a smectite surface
give results comparable with those from quantum mechanical calculations
of the positive charge distribution on isolated solvated cations." On the
basis of this comparison, one would conclude that the increased acidity in
adsorbed water has little to do with the charge distribution on the siloxane
surface (e.g., smectite versus vermiculite) but instead is a local effect of the
presence of exchangeable cations whose charge is screened by solvating
water molecules.

NOTES
1. D. Eisenberg and W. Kauzmann, The Structure and Properties of Water,

Chap. 4. Oxford Univ. Press, New York, 1969. S. A. Rice, Conjectures on the
structures of amorphous solid and liquid water, Topics Curro Chem. 60: 109
(1975). F. H. Stillinger, Water revisited, Science 209:451 (1980).
2. These aspects of the V structure of liquid water are discussed in detail in
F. Hirata and P. J. Rossky, A realization of the "V structure" in liquid water,
J. Chem. Phys, 74:6H67 (IlJHI).

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

73

3. For a discussion of IR and Raman spectroscopy as applied to liquid water, see


W.A.P. Luck, Structure of Water and Aqueous Solutions, Chaps. III and IV.
Verlag Chemie GmBH, Weinheim, Germany, 1974.
4. A good introductory discussion of ESR methods is given in P. L. Hall, The
application of electron spin resonance spectroscopy to studies of clay minerals,
Clay Minerals IS: 321 (1980).
5. S. W. Lovesey and T. Springer, Dynamics of Solids and Liquids by Neutron
Scattering. Springer-Verlag, New York, 1977.
6. D. K. Ross and P. L. Hall, Neutron scattering methods of investigating clay
systems, in Advanced Chemical Methods for Soil and Clay Minerals Research
(J. W. Stucki and W. L. Banwart, eds.). Reidel, Boston, 1980.
7. A careful discussion of the operational meaning of a self-diffusion coefficient is
given in R. Mills, Self-diffusion in normal and heavy water in the range 1 to 45,
J. Phys. Chem. 77: 685 (1973).
8. J. J. Fripiat, The application of NMR to the study of clay minerals, in J. W.
Stucki and W. L. Banwart, op. cit.6
9. G. Sposito, Single-particle motions in liquid water. II: The hydrodynamic
model, J. Chem. Phys. 74: 6943 (1981).
10. For reviews of diffraction methods applied to liquid water, see Chaps. 8 and 9
in Vol. 1 of Water: A Comprehensive Treatise (F. Franks, ed.) Plenum Press,
New York, 1972.
11. D. W. Wood, Computer simulation of water and aqueous solutions, in Water:
A Comprehensive Treatise, Vol. 6 (F. Franks, ed.). Plenum Press, New York,
1979.
12. A. Rahman and F. H. Stillinger, Molecular dynamics study of liquid water,
J. Chem. Phys. 55: 3336 (1971). F. H. Stillinger and A. Rahman, Improved
simulation of liquid water by molecular dynamics, J. Chem. Phys. 60: 1545
(1974).
13. W. L. Jorgensen, Monte Carlo results for hydrogen bond distributions in liquid
water, Chem. Phys. Lett. 70: 326 (1980).
14. See, e.g., Chaps. 8 and 9 in Vol. 1 of F. Franks, op. cit., 10 and Chap. V of
W.A.P. Luck, op. cit.3
15. H. S. Frank and W.-Y. Wen, Structural aspects of ion-solvent interactions in
aqueous solutions: A suggested picture of water structure, Disc. Faraday Soc.
24: 133 (1957). The earliest studies of ion hydration in relation to the
Frank-Wen model have been summarized in B. E. Conway, Ionic Hydration in
Chemistry and Biophysics, Elsevier, Amsterdam, 1981. More recent work is
reviewed in J. E. Enderby and G. W. Neilson, The structure of electrolyte
solutions, Rep. Prog. Phys. 44: 593 (1981).
16. For a discussion of these experiments and their results, see Chaps. 7 and 8 in
Vol. 3 of Water: A Comprehensive Treatise (F. Franks, ed.). Plenum Press,
New York, 1973. A brief summary is given by J. E. Enderby and G. W.
Neilson, op. cit.,15 pp. 647-649.
17. J. E. Enderby and G. W. Neilson, X-ray and neutron scattering by aqueous
solutions of electrolytes, in Water: A Comprehensive Treatise, Vol. 6
(F. Franks, ed.). Plenum Press, New York, 1979. See also A. H. Narten and
R. L. Hahn, Direct determination of ionic solvation from neutron diffraction,
Science 217: 1249 (1982).
IH. N. A. Hewish, J. E. Enderby, and W. S. Howells. Second zone in ionic
solutions. Phys. Rev. Lett. 48: 75fl (19H2).

74

THE SURFACE CHEMISTRY OF SOilS

19. See, e.g., M. Mezei and D. L. Beveridge, Monte Carlo studies ofthe structure
of dilute aqueous solutions of Li+ , Na +, K +, F-, and Cl" ,J. Chern. Phys. 74:
6902 (1981).
20. M. Rao and B. J. Berne, Molecular dynamic simulation of the structure of
water in the vicinity of a solvated ion, J. Phys. Chern. 85: 1498 (1981).
J. Chandrasekhar and W. L. Jorgensen, The nature of dilute solutions of
sodium ion in water, methanol, and tetrahydrofuran, J. Chern. Phys. 77: 5080
(1982).
21. C. H. Lim, M. L. Jackson, R. D. Koons, and P. A. Helmke, Kaolins: Sources
of differences in cation-exchange capacities and cesium retention, Clays and
Clay Minerals 28: 223 (1980).
22. The structural characteristics of halloysite are discussed in detail in the first
three chapters of G. W. Brindley and G. Brown, Crystal Structures of Clay
Minerals and Their X-ray Identification. Mineralogical Society, London, 1980.
23. S. B. Hendricks, On the crystal structure of the clay minerals: Dickite,
halloysite, and hydrated halloysite, Am. Miner. 23:295 (1938). S. B. Hendricks and M. E. Jefferson, Structure of kaolin and talc-pyrophyllite hydrates
and their bearing on water sorption of the clays, Am. Miner. 23: 863 (1938).
24. S. Yariv and S. Shoval, The nature of the interaction between water molecules
and kaolin-like layers in hydrated halloysite, Clays and Clay Minerals 23: 473
(1975). M. I. Cruz, M. Letellier, and J. J. Fripiat, NMR study of adsorbed
water. II: Molecular motions in the monolayer hydrate of halloysite, J. Chern.
Phys. 69: 2018 (1978).
25. P. G. Hall and M. A. Rose, Dielectric properties of water adsorbed by
kaolinite clays, J. C.S. Faraday 174: 1221 (1978).
26. Heat capacity data for halloysite have been reported by M. I. Cruz et aI.,
op. cit. 24 and by P. M. Costanzo, R. F. Giese Jr., M. Lipsicas, and C. Straley,
Nature 296: 549 (1982).
27. J. J. Jurinak, Multilayer adsorption of water by kaolinite, Soil Sci. Soc. Am. J.
27: 270 (1963).
28. J. J. Jurinak and D. H. Volman, Cation hydration effects on the thermodynamics of water adsorption by kaolinite, J. Phys. Chern. 65: 1853 (1961). R. A.
Kohl, J. W. Cary, and S. A. Taylor, On the interaction of water with a
Li-kaolinite surface, J. Colloid Sci. 19:699 (1964). See also J. Fripiat,
J. Cases, M. Francois, and M. Letellier, Thermodynamic and microdynamic
behavior of water in clay suspensions and gels, J. Colloid Interface Sci. 89: 378
(1982).
29. The molecular structure of vermiculite group minerals is discussed comprehensively by G. F. Walker, Vermiculites, in Soil Components, Vol. 2 (J. E.
Gieseking, ed.). Springer-Verlag, New York, 1953.
30. J. F. Alcover, L. Gatineau, and J. Mering, Exchangeable cation distribution
in nickel- and magnesium-vermiculites, Clays and Clay Minerals 21: 131 (1973).
M. I. Telleria, P. G. Slade, and E. W. Radoslovich, X-ray study of the
interlayer region of a barium-vermiculite, Clays and Clay Minerals 25: 119
(1977). J. F. Alcover and L. Gatineau, Structure de l'espace interlamellaire de
la vermiculite Mg bicouche, Clay Minerals 15: 25 (1980). J. F. Alcover and
L. Gatineau, Facteurs determinant la structure de la couche interlamellaire des
vermiculites saturees par des cations divalents, Clay Minerals 15:239 (1980).
J. A. Rausell-Colom, M. Fernandez, J. M. Serratosa, J. F. Alcover, and
L. Gatincau, Organisation de l'espace lnterlumeilaire dans les vermiculites

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

75

monocouches et anhydres, Clay Minerals 15: 37 (1980). V. Fornes, C. de la


Calle, H. Suquet, and H. Pezerat, Etude de la couche interfoliaire des
hydrates 11 deux couches des vermiculites calcique et magnesienne, Clay
Minerals 15: 399 (1980).
31. J. E. Enderby and G. W. Neilson, op. cit.,15 pp. 626-630.
32. C. de la Calle, H. Suquet, and H. Pezerat, Glissement de feuillets accompagnant certains echanges cationiques dans les monocristaux de vermiculites, Bull.
Groupe Fr. Argiles 27: 31 (1975). R. F. Giese and J. J. Fripiat, Water molecule
positions, orientations, and motions in the dihydrates of Mg and Na vermiculites, J. Colloid Interface Sci. 71: 441 (1979).
33. V. C. Farmer and J. D. Russell, Interlayer complexes in layer silicates, Trans.
Faraday Soc. 67: 2737 (1971).
34. H.D.B. Jenkins and P. Hartman, Calculations on a model intercalate containing a single layer of water molecules: A study of potassium vermiculite, Phil.
Trans. Royal Soc. (London) A304: 397 (1982).
35. P. L. Hall, Neutron scattering techniques for the study of clay minerals, in
Advanced Techniques for Clay Mineral Analysis (1. J. Fripiat, ed.), Elsevier,
Amsterdam, 1982. S. Olejnik, G. C. Stirling, and J. W. White, Neutron
scattering studies of hydrated layer silicates, Spec. Disc. Faraday Soc. 1:194
(1970).
36. P. H. Nye, Diffusion of ions and uncharged solutes in soils and soil clays,
Advan. Argon. 31: 225 (1979).
37. D. M. Clementz, T. J. Pinnavaia, and M. M. Mortland, Stereochemistry of
hydrated copper (II) ions on interlamellar surfaces of layer silicates: An
electron spin resonance study. J. Phys. Chern. 72:196 (1973). J. Hougardy,
W.E.E. Stone, and J. J. Fripiat, NMR study of adsorbed water. I: Molecular
orientation and protonic motions in the two-layer hydrate of a Na vermiculite,
J. Chern. Phys. 64:3840 (1976). J. P. Hougardy, P. Tougne, D. Bonnin, and
A. P. Legrand, Study of adsorbed water: Electric potential calculation and
molecular orientation in the two-layer hydrate of a Mg vermiculite, J. Chern.
Phys. 67:5252 (1977). J. J. Fripiat, Organisation des molecules d'eau dans les
silicates de grande surface specifique, Bull. Mineral. 103:440 (1980).
38. G. D. Boss and E. O. Stejskal, Restricted, anisotropic diffusion and anisotropic nuclear spin relaxation of protons in hydrated vermiculite crystals,
J. Colloid Interface Sci. 26: 271 (1968). S. Olejnik and J. W. White, Thin
layers of water in vermiculites and montmorillonites: Modification of water
diffusion, Nature Phys. Sci. 236: 15 (1972). J. Hougardy, J. M. Serratosa,
W. Stone, and H. van Olphen, Interlayer water in vermiculite: Thermodynamic properties, packing density, nuclear pulse resonance, and infrared absorption, Spec. Disc. Faraday Soc. 1: 187 (1970).
39. G. Sposito and R. Prost, Structure of water adsorbed on smectites, Chern.
Rev. 82: 553 (1982). See also. J. Fripiat et ai, op. cit.28
40. This structural arrangement was first proposed in J. Mamy, Recherches sur
l'hydratation de la montmorillonite: Proprietes dielectriques et structure du
film d'eau, Ann. Agron. 19:175 (1968).
41. H. Pezerat and J. Mering, Recherches sur la position des cations echangeables
et de l'eau dans les montmorillonites, Compt. Rend. Acad. Sci (Paris) 265: 529
(1967). R. K. Hawkins and P. A. Egelstaff, Interfacial water structure in
montmorillonite from neutron diffraction experiments, Clays and Gay Minerals 28: 19 (19XO).

76

THE SURFACE CHEMISTRY OF SOilS

42. I. Eger, M. I. Cruz-Cumplido, and J. J. Fripiat, Quelques donnees sur la


capacite calorifique et les proprietes de l'eau dans divers systemes poreux,
Clay Minerals 14:161 (1979).
43. R. Prost, Etude de l'hydratation des argiles: Interactions eau-mineral et
mecanisme de la retention de l'eau. B: Etude d'une smectite (hectorite),
Ann. Agron. 26:463 (1975).
44. D. J. Cebula, R. K. Thomas, and J. W. White, Diffusion of water in Limontmorillonite studied by quasielastic neutron scattering, Clays and Clay
Minerals 29: 241 (1981).
45. R. Calvet, Dielectric properties of montmorillonites saturated by bivalent
cations, Clays and Clay Minerals 23: 257 (1975).
46. P. F. Low, Nature and properties of water in montmorillonite-water systems,
Soil Sci. Soc. Am. J. 43: 651 (1979). J. L. Oliphant and P. F. Low, The relative
partial specific enthalpy of water in montmorillonite-water systems and its
relation to the swelling of these systems, J. Colloid Interface Sci. 89: 366
(1982).
47. See, e.g., D.M.C. MacEwan and M. J. Wilson, Interlayer and intercalation
compounds of clay minerals, in G. W. Brindley and G. Brown, op cit. 22 That
1: 1 electrolytes cannot be dissolved completely in adsorbed water on smectites
and illitic micas has been shown by A. M. Posner and J. P. Quirk, The
adsorption of water from concentrated electrolyte solutions by montmorillonite and illite, Proc. Royal Soc. (London) 278A:35 (1964).
48. See, e.g., Chap. 8 in C. W. Davies, Ion Association (Butterworths, London,
1962) and Chap. 1 in G. H. Bolt, Soil Chemistry. B: Physico-Chemical Models
(Elsevier, Amsterdam, 1979).
49. See Chap. 9 in J. Burgess, Metal Ions in Solution (Ellis Horwood, Chichester,
U.K., 1978) for a discussion of these trends.
50. For an introduction to Lewis acids and the Misono softness parameter, see,
e.g., Chap. 3 in G. Sposito, The Thermodynamics of Soil Solutions. Clarendon
Press, Oxford, 1981.
51. M. M. Mortland and K. V. Raman, Surface acidity of smectites in relation to
hydration, exchangeable cation, and structure, Clays and Clay Minerals 16: 393
(1968). See also J. D. Russell, Infrared study of the reactions of ammonia with
montmorillonite and saponite, Trans. Faraday Soc. 61: 2284 (1965), and M. M.
Mortland, Protonation of compounds at clay mineral surfaces, Trans. 9th Int.
Congo Soil Sci. (Adelaide) I: 691 (1968).
52. J. J. Fripiat, The NMR study of proton exchange between adsorbed species
and oxides and silicate surfaces, in Magnetic Resonance in Colloid and Interface
Science (H. A. Resing and C. G. Wade, eds.). American Chemical Society,
Washington, D.C., 1976. J. Hougardy et al., op. cit. 3?; R. Touillaux,
P. Salvador, C. Vandermeersche, and J. J. Fripiat, Study of water layers
adsorbed on Na- and Ca-montmorillonite by the pulsed nuclear magnetic
resonance technique, Israel J. Chem. 6: 337 (1968). C. Poinsignon, J. M.
Cases, and J. J. Fripiat, Electrical polarization of water molecules adsorbed by
smectites: An infrared study, J. Phys. Chem. 82: 1855 (1978). J. J. Fripiat,
A. Jelli, G. Poncelet, an J. Andre, Thermodynamic properties of adsorbed
water molecules and electrical conduction in montmorillonites and silicas,
J. Phys. Chem. 69:2185 (1965).
53. Chapter 4 in D. Eisenberg and W. Kauzmann, op. cit. I
"4 C Poi
.
I op, cit.'
. ~2
.J..
omsrgnon
et 1\.,

~'

THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES

77

FOR FURTHER READING

:X-

G. W. Brindley and G. Brown, Crystal Structures of Clay Minerals and Their X-ray
Identification. Mineralogical Society, London, 1980. Chapter 3 of this standard
reference contains an excellent discussion of X-ray diffraction studies of adsorbed
water on phyllosilicates.
D. Eisenberg and W. Kauzmann, The Structure and Properties of Water. Oxford
University Press, New York, 1969. This book remains the best short introduction to
the properties of water in all of its phases. Chapter 4 should be read as background
for Sec. 2.1 of the present chapter.
F. Franks. Water: A Comprehensive Treatise, Plenum Press, New York, 19721981. The seven volumes of this encyclopedic reference work that have appeared
thus far contain discussions of all aspects of the chemistry and physics of liquid
water and aqueous solutions. Of special interest are the chapters on bulk liquid
water (Vol. 1), on water in electrolyte solutions (Vols. 3 and 6), and on clay-water
systems, (Vol. 5).
D. J. Greenland and M.H.B. Hayes, The Chemistry of Soil Constituents. Wiley,
Chichester, U.K., 1978. Chapter 6 of this outstanding compendium, written by
V. C. Farmer, reviews infrared spectroscopic studies of adsorbed water.
J. W. Stucki and W. L. Banwart, Advanced Chemical Methods for Soil and Clay
Minerals Research. Reidel, Dordrecht, The Netherlands, 1980. This book provides
an excellent introduction to the use of NMR and INS techniques for the
investigation of adsorbed water structure. Many experimental results for adsorbed
water on phyllosilicates are presented.
J. Texter, K. Klier, and A. C. Zettlemoyer, Water at surfaces, Prog. Surface
Membrane Sci. 12: 327 (1978). This review gives an account of the available data
concerning the properties of water on oxide and organic solid surfaces. The general
conclusions drawn are similar to those stated in the present chapter for the
structure of adsorbed water on phyllosilicates, except that hydrogen bonding of the
water to the surface plays a more prominent role.

3
THE ELECTRIFIED INTERFACE
IN SOILS

3.1. THE BALANCE OF SURFACE CHARGE

The creation of an interface between a soil solution and the solid phases of
a soil clay induces, by definition, fundamental dissymmetries in the
molecular environment of the interfacial region. The forces acting on a
cation bound into a siloxane ditrigonal cavity on a dry smectite surface, for
example, are entirely different from those acting on the same cation when
it is bathed by an aqueous solution phase, and the behavior of a cation
immersed in the soil solution near a smectite surface is very different from
its behavior near a small anion in the bulk aqueous phase. The net effect of
the molecular constituents of one phase on those of an adjacent phase is a
structural reorganization at the interface that reflects a compromise among
competing interactions originating in the bulk phases. It is, of course, for
this reason that the structure of liquid water near a solid phase is different
from that in the bulk liquid and the distribution of charge in both the solid
and aqueous solution phases near an interface becomes distorted from
what exists in bulk matter. Cations bound into inner-sphere surface
complexes may solvate and thereby become farther displaced from the site
of the negative charge they balance. Ions in the soil solution can respond
differently to an altered water structure or to shifts in the charge distribution in the solid phase at the interface by assuming configurations that are
not electrically neutral in a representative volume element in the aqueous
phase. These perturbations of the molecular environment lead intrinsically
to persistent separation of charge and therefore to an electrified solid-liquid
interface in soils.
The most important physical characteristic of an electrified interface is
its surface charge density. The concept of surface charge density was
introduced in Sec. 1.5 in the discussion of the surface density of intrinsic,
permanent structural, and net proton charge. These three surface charge
densities are related by the equation
(J.I)

THE ELECTRIFIED INTERFACE IN SOILS

79

where (Tin is the intrinsic surface charge density, (To the permanent
structural surface charge density, and (TH the net proton surface charge
density. Each term in Eq. 3.1 can be measured either in coulombs per
square meter or in moles of charge per square meter, and each can be
either positive or negative.
Besides the intrinsic surface charge density, two other components of the
density of surface charge on a soil particle can be defined. The surface
density of inner-sphere complex charge, (TIS, is equal to the net total surface
charge of the ions, other than H+ or OH- , that have formed inner-sphere
complexes with the surface functional groups in a soil. Examples of these
complexes were given in Figs. 1.8 and 1.9, where surface complexes
between vermiculite and K+ and between goethite and HPO~- are
illustrated. Other examples include the complex between Pb2+ and the
hydroxyl groups on alumina and that between Fe3+ and the carboxyl
groups on soil organic matter. 1 The generic term specific adsorption is
often used to describe the effects of inner-sphere surface complexation of
ions in the soil solution by surface functional groups on soil clays.
The surface density of outer-sphere complex charge, (Tos, is equal to the
net total surface charge of the ions that have formed outer-sphere
complexes with the surface functional groups in a soil. Examples of these
complexes are found in Figs. 1.8 and 1.10, where surface complexes
between Ca 2+ and montmorillonite and between Na + and kaolinite are
shown. Other typical examples are the complex between 0- and protonated aluminol groups and that between Mn2+ and carboxyl groups on soil
organic matter. 1 The generic term nonspecific adsorption can be applied to
outer-sphere surface complexation of ions by the functional groups exposed on soil clay particles.
With these additional definitions, the surface density of net total particle
charge can be expressed mathematically:
(Tp => (Tin

= (TO

(TIS
(TH

(TOS

(TIS

(Tos

(3.2)

Each of the terms on the right side of Eq. 3.2 can be either positive or
negative, but in general their sum will not equal zero despite the possibility
for cancellation. The balance of surface charge, as implied above, cannot
be expected to hold, in general, for only part of the interfacial region.
What is yet missing is the equivalent surface density of dissociated charge,
(To. This quantity is equal to minus the net total particle charge neutralized
by ions in the soil solution that have not formed complexes with surface
functional groups. These ions, whether positive or negative, are fully
dissociated from the surfaces of the solid particles in a soil and are free to
move about in the soil solution beyond the interfacial region. The balance
of surface charge can now be expressed by a combination of Eq. 3.2 and
(Tn:

('0

t-

('11

('" +
+ (T,S

('n = ()

(T( IS

+ (TI I

=: ()

(3.3a)
(3.3h)

80

THE SURFACE CHEMISTRY OF SOILS

Equation 3.3 is the fundamental conservation law that must be satisfied by


the electrified interfaces in any soil.
The common methods for measuring 0"0 and O"H are outlined in Sec. 1.5.
Conventional methods for determining the remaining densities in Eq. 3.3
have not been established, but there are several techniques that are useful
in particular circumstances:
1. If O"in has been measured for an assembly of soil clay particles, then the
sum O"IS + O"os + O"D can be calculated with the help of Eq. 3.3.
2. If it is assumed that the cation and anion exchange capacities (CEC and
AEC) of a soil clay refer only to those ions in the interfacial region that
can be displaced easily by a leaching solution, then the difference
CEC - AEC is proportional to the sum O"os + O"D' In mathematical
terms,
F(CEC - AEC)
(3.4)
O"os + O"D =
S
where F is the Faraday constant, S is the specific surface area of the soil
clay, and CEC and AEC are expressed in moles of charge per unit mass
of soil clay. Equation 3.4 expresses the concept that AEC - CEC is
proportional to the surface density of intrinsic charge after correction
for any inner-sphere complex charge. This concept reflects the wellknown low desorbability of specifically adsorbed ions.
3. If O"in is known and O"os + O"D is measured by the method described in
item 2, then O"IS can be calculated by rearranging Eq. 3.3:
O"IS = -O"in +

F(AEC - CEC)

(3.5)

The determination of O"in is carried out in the absence of specific


adsorption, but otherwise under the same soil conditions as exist for the
determination of CEC and AEC in the presence of specific adsorption.
4. If it is assumed that the electrokinetic plane of shear near a soil particle
coincides with the outer periphery of its surface complexes, then
electrokinetic mobility experiments/ can be interpreted to provide an
estimate of O"p and, by Eq. 3.3a, of O"D' The theoretical basis for this
method is discussed in Sec. 3.4. It may be noted in passing that no
assumptions about the detailed structure of the interfacial region are
required in order to measure a zero value for O"D' Given the single
assumption about the plane of shear, O"D vanishes at zero electrokinetic
mobility.
The parameters O"IS' O"os, and O"D reflect the disposition of ions in the soil
solution after they have become incorporated into the interfacial region.
Therefore, these surface charge densities represent the net charging effects
of the surface speciation of the ions. By analogy with the use of speciation
models (ion-association models) to estimate the distribution of ionic charge
in aqueous phases like soil solutions;' surface speciation models (surface

THE ELECTRIFIED INTERFACE IN SOILS

81

complexation models) could, in principle, be used to estimate (TIS' (Tos,


and (Tn individually. The basic properties of surface complexation models
are discussed in Chap. 5. Suffice it to say here that these models have been
developed well enough to permit quantitative, verifiable estimates of
surface charge densities under reasonable physical assumptions.
3.2 POINTS OF ZERO CHARGE
Points of zero charge are Elf values associated with specific qHUliJ\QOs
iIllPosed on one or more of the. sqt:fac~ cha.t,gePensiti~s d.es9ri,1?~4jn
Sec. 3.1. The definitions ofthe most important points ofzero charge in the
mrface chemistry of soils are summarized in Table 3.1.
The conventional e~inl~qL;Z!?!2~c:~q,~G!,!R~~2..~~..t~~'4:,I;ty~lu~. 2L ~~~911..
solution when the.t()t::l1g.~tpartiFle cl:1arg~.YflIli~hc;:s. By Eq. 3.3a, this
condition is met when '!n ~O. The PZC. ca~ .be. measured directl.rjn
electrokineticexperimentsalld" inconoi~'ar~sfllo~fif"si~~Tes-;Ss ince-both
invo'lve"phenomeiia sensitive to the"total 'net diarge oo'sU'spended particles.
The point of zero net proton charge (PZNPC) is the pH value of the soil
solution at which (TH, defined in Eq. 1.24, is equal to zero. As can be
inferred from Eq. 1.25, the PZNPC can be measured by potentiometric
titration, provided only proton-selective surface functional groups on the
soil solids are titrated.
The point of zero salt effect (PZSE), defined in Eq. 1.27, also can be
measured by potentiometric titration. The PZSE is determined by locating
the common point of intersection of several graphs of (TH versus pH, each
determined at fixed ionic strength of the electrolyte background. Another
common method for measuring the PZSE involves the equilibration of a
set of soil suspensions, initially adjusted to a range of pH values expected
to include the PZSE, with a background electrolyte at two different ionic
strengths. After the suspensions have come to equilibrium, the pH values
of their aqueous solution phases are determined, and the pH value that
shows no change with ionic strength is designated as the PZSE. 6
The point of zero net charge (PZNC) is the pH value of a soil solution at
which the difference CEC - AEC equals zero. This difference is proportional to either an optimal or a nonoptimal value of the intrinsic surface
Table 3.1. Definitions of some points of zero charge

Name

Symbol
PZC a
PZNPC
PZSE
PZNC

Point
Point
Point
Point

of
of
of
of

zero
zero
zero
zero

charge
net proton charge
salt effect
net charge

Defining equation
= 0
UH = 0
(auH/alh = 0
UD

Uos

AINU termed IEI', iNlleleclrk pulnl, when measured by an electroklnetlc experiment.

UD

=0

THE SURFACE CHEMISTRY OF SOILS

82

charge density, as pointed out in Sec. 1.5. Therefore, the PZNC can be
measured by the Schofield method" applied over a range of pH values. If it
is assumed that the reactant salt solution used to saturate the soil with the
two index ions can displace only the ions contributing to aos and aD, then
a nonoptimal value of ain is measured and the PZNC corresponds to the
condition aos + aD = 0 (Table 3.1). Otherwise, if it is known that the
reactant salt solution can displace even specifically adsorbed ions, then it is
appropriate to write
ars + aos + aD =

F(CEC - AEC)
S

(3.6)

instead of Eq. 3.4, and the optimal value of ain is measured. In this case,
the PZNC corresponds to the condition ain = O. The commonly measured
points of zero charge are illustrated in Fig. 3.1. 7
It is evident from Table 3.1 that the surface charge density conditions
that define points of zero charge are not the same and therefore that there
can be numerical differences among these pH values for the same soil
particles. The circumstances that permit equality among the points of
zero charge can be ascertained directly, however, through an appeal to
the charge conservation law in Eq. 3.3. Consider as a simple case the
possibility of equality between the PZC and the PZNPC. According to
Table 3.1 and Eq. 3.3b, this possibility is realized when the equation
ao

+ ars +

aos = 0

(3.7)

is valid. This charge balance equation can be satisfied in an infinitude of


ways, one of which is the independent vanishing of each of the three
component surface charge densities. Although this special case is quite
unlikely in soils, it can be achieved approximately for reference soil
minerals suspended in aqueous solutions of 1 : 1 electrolytes, as exemplified
by the data for -y-A1 203 , birnessite, and corundum in the second and third
columns of Table 3.2. On the other hand, Eq. 3.7 does not appear to apply
to a comparison of PZC with PZNPC for goethite (unless other factors,
such as the method of solid preparation, surface impurities, and crystallinity, are operating).
If the PZC and the PZNC are to be equal and if Eq. 3.4 is assumed
correct, then Table 3.1 and Eq. 3.3b demand that aos vanish identically.
For, if a soil solution is at the PZC, then aD = 0 and the condition for the
PZNC (aos + aD = 0) requires that aos = 0 also. Conversely, if a soil is
at the PZNC, then aos = - aD by hypothesis and the soil is at the PZC as
well only if aos = O. To determine whether aos = 0 at either the PZC or
the PZNC, one would need to speciate the adsorbed ions in a soil into
outer-sphere surface complexes versus completely dissociated species. This
could be done, for example, by studying the electrokinetic or coagulation
behavior of soil particles that have been brought to the PZNC. On the
molecular level, equality between the PZC and the PZNC is expected if
thc soil particles are suspended in a I : I electrolyte solution wherein both

THE ELECTRIFIED INTERFACE IN SOILS

83

>
rVI

c>

.:<:

(\J

ex>

0
><

pH

-I

::l

-2

GOETHITE
0.015 M NaCI04
BAR-YOSEF et al. (1975)

Y- AI203
10-3M NaCI
HUANG AND STUMM (1973)
4

-60

rc>

-20

40
KAOLINITE
FERRIS AND JEPSON (1975)

.:<:
u

0
E

c>

0"

.:<:

HYDROXYAPATITE
BELL et ol. (l973l

>-

Z
0"

pH

KCI
'"
o

1M
O.IM

O.OIM

POINTS OF ZERO CHARGE

11.0

Figure 3.1. Experimental examples of the PZC of y-Alz0 3 , the PZNPC of


goethite, the PZSE of hydroxyapatite, and the PZNC of kaolinite." (u is the
electrophoretic mobility.)

the cation and the anion form only outer-sphere surface complexes. This
condition appears to be met approximately for birnessite and kaolinite,
according to the data in the second and fifth columns of Table 3.2.
Equality between the PZC and the PZSE obtains if the equation
(O'IS

O'OS)11

(O'IS

0'0S)12

(3.8)

holds, where 11 and 12 refer to two different ionic strengths of the soil
solution. Equation 3.R is derived by applying Eq. 3.3b at each ionic
strength. noting the definitions in Table 3.1, and deleting Un because

THE SURFACE CHEMISTRY OF SOILS

84

Table 3.2. Comparison of points of zero charge for several solid phases suspended
in solutions of 1 : 1 electrolytes
Solid

PZC

PZNPC

PZSE

Alon (y-Al z03)


Birnessite (8-MnOz)
Calcite (CaC0 3)
Corundum (a-Alz03)
Goethite (a-FeOOH)
Hematite (a-FeZ03)
Hydroxyapatite (Cas(P04hOH)
Kaolinite (Si4AI4OlO(OHs))
Quartz (a-SiO z)

8.7

8.2 0.5
2.2 0.7

8.5

1.7
10
9.1
6.1

0.4
1
0.2
0.6

7.5 0.1
4.7
2.0 0.3

2.3 1.1
9.5

PZNC
1.9 0.5

9.1
7.7 0.2
8.4 0.1

7.3 0.2
8.5
7.6 0.2
4.8
2.9 0.9

Sources of data: Alon, c.-P. Huang and W. Stumm, J. Colloid Interface Sci. 43:409 (1973); S. Goldberg,
personal communication. Birnessite, L. S. Balistrieri and J. W. Murray, Geochim. Cosmochim. Acta
46: 1041 (1982); R. M. McKenzie, Aust. J. Soil Res. 19:41 (1981). Calcite, G. A. Parks, in Chemical
Oceanography, Academic Press, London, 1975, Vol. Y, pp. 241-308; R. J. Hunter, Zeta Potential in
Colloid Science, Academic Press, New York, 1981, pp. 228ft. Corundum, S. Goldberg, personal
communication; R. J. Hunter, op. cit. Goethite, G. A. Parks, op. cit.; S. Goldberg, personal communication; T. L. Theis and R. O. Richter, Advan. Chem. Series 189: 73 (1980); L. S. Balistrieri and J. W.
Murray, Am. J. Sci. 281:788 (1981). Hematite, S. Goldberg, personal communication; A. Breeuwsma and
J. Lyklema, J. Colloid Interface Sci. 43:437 (1973). Hydroxyapatite, Bell et aI., J. Colloid Interface Sci.
42:250 (1973). Kaolinite, J. Baham, personal communication; A. P. Ferris and W. B. Jepson, J. Colloid
Interface Sci. 51:245 (1975). Quartz, R. J. Hunter, op. cit.

permanent structural charge cannot be affected by changes in ionic


strength. It follows from this result that the PZC is the same as the PZSE if
no surface complexes form. Although this condition is sufficient, it is not
likely to occur in soils. A better assumption is the lack of inner-sphere
surface complexes if soil particles are suspended in a 1: 1 electrolyte
solution. In that case, Eq. 3.8 reduces to
(3.9)
which implies that the surface density of outer-sphere complexes remains
invariant under a change in ionic strength. This condition can be met if Uos
is at its maximum value at both 11 and 12or if the cation and the anion have
about the same affinity for the soil particles and adsorb or desorb equally
with changes in ionic strength. Evidently, ions such as Na+ and Cl", which
are often used in background electrolytes, meet the criterion of roughly
equal affinity for oxide surfaces, as exemplified by birnessite and quartz in
Table 3.2. (Goethite appears to be an exception.)
If the background electrolyte contains ions that can form inner-sphere
surface complexes (e.g., SO~- or Ca 2+), then Eq. 3.8 must be retained as
the general condition for equality between the PZC and the PZSE. The
condition applied to Urs separately can be met through a change in ionic
strength while the concentration of the specifically adsorbing ion is held
constant. For example. the concentration of SO~- could be maintained
fixed while: the ionic strength is changed through a shift in the concentra-

THE ELECTRIFIED INTERFACE IN SOILS

85

tion of a swamping electrolyte, such as NaCl. In this case multiple plots


of (TH versus pH at fixed ionic strength intersect at a common point that
determines the PZC as well as the PZSE. However, at any ionic strength,
Eq. 3.3b requires the condition
(pH =

pzq

(3.10)

to hold at the PZC. If (TIS < 0 (specific anion adsorption), the value of (TH
will be larger than when no specific adsorption occurs and, since (TH
increases with proton activity, the PZC must shift downward relative to
that case. If (TIS> 0 (specific cation adsorption), the value of (TH at the
PZC will be smaller than when no specific adsorption occurs and the PZC
must shift upward. This kind of shift in the PZC is observed commonly in
soils. An example of the case (TIS < 0 (o-phosphate adsorption) is shown in
Fig. 3.2. 8
Should the swamping electrolyte solution contain an ion that can adsorb
specifically, matters change considerably as regards equality between the
PZC and the PZSE. In this case, Eq, 3.8 cannot hold unless (TIS has
reached a maximum value that remains invariant under a change in the
concentration of the specifically adsorbing ion. Otherwise, there must be a
different value of the PZC at every ionic strength of the swamping
electrolyte, and points of intersection of graphs of (TH versus pH cannot be
used to determine the PZC. If (TIS < 0 (specific anion adsorption),
Eq. 3.10 implies that (TH increases with increasing ionic strength (which
decreases (TIS) and the PZSE shifts upward from the PZC determined
when (TIS = O. If (TIS> 0 (specific cation adsorption), the same line of
reasoning implies a downward shift of the PZSE. Note that these shifts are
opposite in sense of what occurs when the concentration of specifically
adsorbing ion is maintained constant while the ionic strength is changed.
This fact has led to some confusion about the expected effects of specific
adsorption on the PZC when the PZSE has been equated incorrectly to the
PZC in a suspension containing a specifically adsorbing, swamping ion. 1 9
Figure 3.2. The downward shift of the PZSE for an Oxisol soil (Typic Torrox) in
response to the specific adsorption of o-phosphate."
MOLOKAI s.C.1.
OM PHOSPHATE

:r

cr
I

:r
cr

MOLOKAI s.c.I,
0.0156 M PHOSPHATE

~
1M
0 0.1 I!!
O.OIM
0 0.001 I!!

THE SURFACE CHEMISTRY OF SOilS

86

It is possible to distill from the present discussion two general rules


concerning equality among the points of zero charge summarized in
Table 3.1. These rules depend only on the applicability of the balance of
surface charge (Eq. 3.3) and not on any detailed model of the interface
between the soil solution and soil solid phases.
1. If a soil is suspended in a swamping electrolyte solution of a 1: 1

electrolyte whose cation and anion form only outer-sphere surface


complexes, the PZC, the PZNC, and the PZSE for the soil are likely to be
equal.
2. If a soil is suspended in a swamping 1: 1 electrolyte as in rule 1, along
with a fixed concentration of an electrolyte containing an ion that adsorbs
specifically, the PZC and the PZSE for the soil are likely to be equal. In
this case, the PZC changes relative to its value as determined by rule 1
and the sign of the change is the same as the sign of the valence of the
specifically adsorbing ion.
The relationship between the PZNPC and the PZNC for a soil when the
experimental conditions of rule 1 are met can be deduced from Eq. 3.3b
and Table 3.1. Consider the PZNPC. Equation 3.3b becomes
(To

(TOS

(TD

= 0

(pH = PZNPC)

(3.11)

Equation 3.11 does not contain (TIS because inner-sphere surface complexes are assumed to be absent, in accordance with rule 1.
At the PZNC, on the other hand, Eq. 3.3b and Table 3.1 lead to the
charge-balance expression
(pH

= PZNC)

If the PZNPC is larger than the PZNC, then

(3.12)

at the PZNC must be


larger than (TH at the PZNPC since (TH increases as the pH decreases. But
(TH (PZNPC) = 0 by hypothesis, and therefore (TH (PZNC) must be a
positive quantity. It follows immediately from Eq. 3.12 that (To must be a
negative quantity. The same kind of reasoning also shows that, if the
PZNPC is smaller than the PZNC, then (To must be a positive quantity.
These conclusions can be stated as the following general rule:
(TH

3. If a soil is suspended in a swamping electrolyte solution of a 1: 1


electrolyte whose cation and anion form only outer-sphere surface
complexes, the sign of the difference PZNPC - PZNC is opposite the sign
of the surface density of structural charge, (To. If PZNPC = PZNC, then
the structural surface charge density must vanish.
l

It should be noted that the relation between the sign of (To and that of
PZNPC - PZNC does not depend on equality between the PZC and the
PZNC but only on the condition that (Trs = o.

THE ELECTRIFIED INTERFACE IN SOILS

87

Suppose now that Eq. 3.4 applies in addition to Eq. 3.11. The combination of these two equations then produces the expression
F(CEC - AEC)

0"0

+=0

(3.13)

valid at the PZNPC. Equation 3.13 shows that a measurement of the


difference between CEC and AEC at the PZNPC can be used to calculate
the structural surface charge density in a soil. Note that the sign of 0"0 is
opposite that of the difference CEC- AEC. In moderately weathered
soils, this difference is expected to be positive and thus 0"0 is negative
because of isomorphic substitutions of cations of lower valence for those of
higher valence in phyllosilicates. In highly weathered soils, the difference
CEC - AEC may be negative at the PZNPC and (To can be positive
because of isomorphic substitutions of cations of higher valence for those
of lower valence in hydrous oxides. 10
Before leaving this discussion of the conceptual basis of the point of zero
charge, two experimental aspects of the measurement of the PZC in a soil
should be mentioned. First, note that the condition on O"H stated in
Eq. 3.10 is impossible to fulfill experimentally if 0"0 is a negative quantity
large enough to make (TH so large at the PZC and the pH value of the soil
solution so low that it cannot be achieved experimentally without dissolving the soil particles. This set of circumstances appears to exist for soil clays
dominated by 2: 1 phyllosilicates, which exhibit large absolute values of 0"0
because of isomorphic substitutions and for which the PZC has not been
measured successfully. 1
A second experimental aspect of the PZC deals with the relationship of
the PZC of a soil to those of its individual mineral constituents." As an
illustration of this point, consider a simple mechanical mixture of two kinds
of mineral particle, A and B. At the PZC, the total particle charge in the
mixture is zero:
(pH

PZC)

(3.14)

where S is the specific surface area and m is the mass of particles of


either kind. After dividing Eq. 3.14 by the total mass of the mixture,
m = m: + mB, one can rearrange the expression to have the form
O"DASA

:A

O"DBSB( 1

:A)

=0

from which it follows that


WA

O"DBSB

----==-=--(TDBnB -

(pH = PZC)

(3.15)

O"DASA

where WA = mAim is the mass fraction of A in the mixture. Equation 3.15


permits the calculation of the PZC of the mixture as a function of W A' If
the specific surface areas of the two component solids are known. and if the

88

THE SURFACE CHEMISTRY OF SOILS

surface charge density, lTo (at fixed ionic strength), has been measured for
each as a function of pH value, then the value of W A corresponding to each
pH value in the measured domain can be calculated with Eq. 3.15. These
pH values can be equated to the PZC corresponding to the calculated W A'
For example, if at pH 6 the product lTOASA is equal to -0.02 mol.kg"! and
lTOBSB is equal to +0.18 mol.kg", then, according to Eq. 3.12, pH 6 is
the PZC of a mixture containing a mass fraction of 0.9 for component
solid A. As a general rule, Eq. 3.12 does not lead to a linear relationship between the PZC and WAY Thus the PZC of a mixture cannot be
predicted by a simple linear interpolation between the PZC of component
B and that of component A as W A increases from 0 to 1. For a soil
comprising several constituent minerals in the clay fraction, no linear
relation between the soil PZC and the PZC of the constituent minerals is
expected.
3.3. POTENTIALS NEAR AN ELECTRIFIED INTERFACE

The development of surface charge at the interface between soil particles


and the soil solution is a reflection of inhomogeneities in the molecular
environment of the interface, as discussed in Sec. 3.1. These molecular
inhomogeneities also influence the thermodynamic properties of both the
charged species in the soil particles and those in the soil solution. In
particular, the distribution of a charged species between the two bulk
phases, regardless of their composition, is determined by the electrochemical potential of that species, ji.. The gradient of the electrochemical
potential of a species drives its diffusive transfer between phases, and
equilibrium with respect to this transfer is described by the equality
(3.16)
where A and IT denote two different phases that contain the charged
chemical species i. The electrochemical potential has the units joules per
mole. 12
The measurement of the electrochemical potential can be illustrated by
considering the following soil clay suspension-soil solution system into
which two identical electrodes are immersed:

Ag;AgCl

NaX(s),NaCI(aq)

R
NaCI(aq)
solution

AgCI;Ag

The single vertical line refers to the interface between a silver-silver


chloride electrode and either a soil suspension or a soil solution. The
double vertical line marked W refers to a membrane that is impermeable to
the soil colloidal anion. X-I, but permeable to dissolved ions and water.
There may be ions other than Na +- and CI- in the suspension or in the
solution, but they are assumed not to interfere with the behavior of the

THE ELECTRIFIED INTERFACE IN SOILS

89

silver-silver chloride electrode toward Cl" anions. At the left electrode


(L), by convention.P the oxidation reaction
Ag(s) + Cl-(L,aq) = AgCI(s) + e
takes place, where L refers to a point inside the left electrode assembly and
e denotes an electron. At the right electrode (R), the reduction reaction
AgCI(s) + e = Ag(s) + Cl-(R,aq)
takes place, where R refers to a point inside the right electrode assembly.
The overall electrode-pair reaction, obtained by combining the two
half-reactions, is
(3.17)
According to the thermodynamic conventions for the assignment of emf
values to electrode assemblies, summarized in Table 3.3, the emf across
the silver-silver chloride electrode pair is given by the equation
(3.18a)
where F is the Faraday constant and E is the emf in volts. It is expected that
equilibrium exists between Cl-(L,aq) and Cl" anions in the suspension as
well as between Cl-(R,aq) and Cl" anions in the solution. Therefore, by
Eq. 3.16,
and
where so refers to the solution and su to the suspension. The combination
of the two parts of Eq. 3.18 leads to an expression for the difference
between the electrochemical potential of the chloride ion in the suspension
and that in the aqueous solution:
(3.19a)
Equation 3.19a illustrates the general rule that differences in electrochemical potential for a charged species can be measured by determining the
Table 3.3. Conventions in the assignment of emf to galvanic cells13
1. The cell reaction is written as if oxidation occurs spontaneously at the left

electrode and reduction at the right electrode.


2. If the overall cell reaction is
aA + bB + ... = xX + yY + ...

the cell emf (in volts) may be defined by the relation


XM[X] + YM[Y] + ... - au. [A] - bM[B] - ...

-FE

where F is the Faraday constant and E is the cell emf.


3. In these conventions. it is assumed that all reduction or oxidation half-reactions
are written in terms of the transfer of 1 mole of electrons.

90

THE SURFACE CHEMISTRY OF SOILS

emf developed across a pair of electrodes that behave reversibly toward the
charged species. Clearly, the derivation of Eq. 3.18 can be carried through
for any charged species in two different aqueous systems by using
electrodes that behave reversibly toward the species. The corresponding
general result for the electrochemical potential difference is

fl;iLi = iLA[i] - iLCT[i]

ZiFE

(3.19b)

where A denotes the phase containing an electrode (reversible to charged


species i) at which a reduction occurs and (J" denotes the phase containing
the electrode at which an oxidation occurs. The parameter Z, is the valence
of species i. If equilibrium exists with respect to the transfer of species i
between the phase A and (J", then Eq. 3.16 applies and E = 0 in Eq. 3.19b.
Thus the absence of an emf across a pair of electrodes that behave
reversibly toward a charged species can be used to indicate equilibrium
with respect to the transfer of the species between two phases.
The electrochemical potential of a charged species can be envisioned as
the potential difference (in the sense of mechanics) involved with the
transfer of 1 mole of a charged species from a point at charge-free infinity
to a point inside a material phase. 12 It is evident from this conceptualization that iLA[i] depends on the purely chemical nature of the species i as
well as on the purely electrostatic interactions that can occur between i and
other charged species in the phase A during the transfer process. For
example, in the case of the chloride ion discussed above, iLsu[CI-] should
depend on the chemical properties of chloride in the suspension as well as
on the electrostatic interactions between Cl" and either the other ions or
the charged solid surfaces in the suspension. Similarly, iLso[CI-] should
depend on the chemical properties of chloride in the soil solution and on
the electrostatic interactions between Cl" and the other dissolved ions.
This dual characteristic of the electrochemical potential suggests that it is
worthwhile to inquire as to the physical significance of the formal
definition:

P-[iJ == go + RT In(i) + ZiFcP

(3.20)

where go is a function of temperature and pressure, as well as of the


"purely chemical" nature of the species whose activity is (i), R is the molar
gas constant, T is the absolute temperature, and cP is the electric potential
to which i is subjected. Evidently Eq. 3.20 would separate the electrochemical potential into a purely chemical part-the first two terms on the
right side-and a purely electrostatic part containing the potential, cPo
Consider now Eq. 3.20 applied to Cl" in the soil clay suspension-soil
solution system diagramed below Eq. 3.16. The left side of Eq. 3.19a can
be expressed

flsO
su JLCI

= /:

so .so

[Cl-] - c. [Cl"] + RT In [(CI-)so] - F(.l.. _.l..)


so .SII
(Cl " )su
'l'so
'l'SIi

p.21)

THE ELECTRIFIED INTERFACE IN SOILS

91

The left side of Eq. 3.21 is well defined and measurable, as indicated in
Eq. 3.19a. The right side of Eq. 3.21 contains the difference between go
for 0- in the two aqueous systems, the ratio of the activities of 0-, and
the Donnan potential difference cPso - cPsu. These three quantities have
physical meaning if it is possible to measure any two of them unambiguously, i.e., without making unverifiable assumptions about the nature of
the two aqueous systems. Unfortunately, no experimental method exists
that can determine even ratios of single-ion activities without making
unverifiable extrathermodynamic assumptions. Moreover, no experimental technique exists for an unambiguous measurement of the difference
in go values for two phases of different chemical composition, and no
electrode assembly can measure a Donnan potential difference without
the data being interpreted through unverifiable extrathermodynamic
assumptions. 12 It follows that, in this case, the partitioning of the right side
of Eq. 3.21 has no physical significance.
Suppose that there is good reason to believe that all of the chloride in the
suspension diagramed below Eq, 3.16 is in dissolved form and therefore
that the Standard State of Cl" is the same in the suspension and the
aqueous solution. In addition, suppose that equilibrium exists with respect
to the transfer of chloride between the two aqueous systems. Then
Eq. 3.21 can be reduced to the expression
cPsu - cPso = -RT In ~msu~
-l'
m so

RT In [Ysu]
+ -l'

Yso

(3.22)

where m is a chloride molality and y is a chloride activity coefficient. 14 The


molality of Cl" in the suspension and in the aqueous solution can be
determined experimentally. Therefore, the physical significance of the
Donnan potential difference hinges on the measurability of the activity
coefficient ratio, Ysu/Yso' This ratio might be estimated with the help of
model equations based in molecular theory, but no experimental technique exists that can determine the ratio without unverifiable extrathermodynamic assumptions.F Therefore, even in this relatively simple
case, the Donnan potential difference remains without empirical meaning.
The electric potential on the right side of Eq. 3.20 is known as an inner
potential, and the potential difference on the right side of Eq. 3.21 is an
example of a Galvani potential difference,
(3.23)
between two phases denoted A and a. The discussion presented in this
section is intended to illustrate the general principle'< that Galvani
potential differences between two phases of different chemical composition
cannot be measured. The root problem is that the inner potential is the
electric potential difference between a point at charge-free infinity and a
point inside a material phase. Although this potential is a well-defined
entity in classical electrostatics (the theory of a hypothetical charged fluid),
it is not well defined in thermodynamics (the macroscopic theory of matter

THE SURFACE CHEMISTRY OF SOilS

92

in stable states) because the interactions that produce the chemical


properties of charged species cannot be divided unambiguously into
electrostatic and nonelectrostatic categories. However, if a charged species
occurs in two phases with identical chemical composition, this difficulty is
obviated by the fact that the purely chemical part of the electrochemical
potential of a charged species must be the same in the two phases. It
follows from Eq. 3.20 that, in this case,
(3.24)
and therefore that Galvani potential differences between two phases of
identical chemical composition are measurable. Examples of the application of Eq. 3.24 include Galvani potential differences between two identical metal wires or between two identical aqueous solutions.F For two
aqueous solutions having the same chemical constituents but different
concentrations of those constituents, Eq. 3.20 leads to the expression
(3.25)
where Aand A' denote the two solutions. If the ratio of activities in the two
solutions can be determined (e.g., by the use of activity coefficients or with
an ion-selective electrode), then, within the conventions prescribed for
interpreting the activity determination, the Galvani potential difference
~;cP becomes a measurable quantity.F
The concepts of electrochemical and inner potential can be used to
classify interfaces, as show in Table 3.4. If charged species cannot traverse
an interface freely, the interface is called polarizable and the condition for
equilibrium across the interface is that zero Galvani potential difference
exists across it. If charged species can traverse the interface freely, it is
called reversible (or nonpolarizable) and the condition for equilibrium
across the interface is that zero electrochemical potential difference exists
across it. Thus a polarizable interface is analogous to a capacitor and a
Table 3.4. The categories of interfaces

Interface
type
Polarizable

Ions traverse
freely?

Equilibrium
condition

Equivalent
circuit

No

Rt
Reversible

Yes

oo

THE ELECTRIFIED INTERFACE IN SOILS

93

resistor of infinite resistance arranged in parallel, whereas a reversible


interface is analogous to a capacitor and a resistor of zero resistance
arranged in parallel. Since the behavior of the reversible interface is
governed by the electrochemical potentials of the charged species that can
cross it freely, these species are called potential-determining for the
interface.
An. example of a polarizable interface is that between a mercury
electrode and liquid water, since the concentration of mercury ions in the
aqueous phase is quite negligible. In this case, it is common to assume as a
practical convention that equilibrium at the interface exists when the emf
of the mercury electrode-reference electrode pair vanishes, since a
Galvani potential difference between mercury and water cannot be measured. An example of a reversible interface is that between a hydrous oxide
solid and liquid water. In this case, H+ and OH- ions can cross the
interface freely and are potential-determining. Equilibrium at the interface
is established when the net ion transport across the interface vanishes, i.e.,
when there is no change in the pH value of the aqueous phase. Note that
the interface between a soil particle and the soil solution is in general
reversible. Any charged species that is adsorbed by the particle and found
in the soil solution is potential-determining.
Although the inner potential has no strict thermodynamic significance, it
can always be defined and studied in the context of a model of the
interfacial region. The use of the inner potential in this way must be
understood clearly to have no physical significance outside the conventions
that establish the model. A case in point is the inner potential 0/ defined in
diffuse double layer (DDL) theory through the Poisson-Boltzmann
equation.P This electric potential was used in Sec. 1.4 to derive a relationship between the exclusion volume for soil clay suspended in a 1: 1
electrolyte solution and the electrolyte concentration (Eq. 1.18). The final
result did not exhibit 0/ and therefore was verifiable by experiment even
though 0/ itself is not measurable. This kind of model definition of the inner
potential is not invalid so long as it is self-consistent. However, in no sense
would it be correct to substitute a solution of the Poisson-Boltzmann
equation into Eq. 3.20 and then claim that c/J had thereby been made
legitimate. This step would be incorrect conceptually because c/J in Eq. 3.20
applies to an entire phase whereas 0/ obtained by solving the PoissonBoltzmann equation is an explicit function of position within a phase. Put
more generally, 0/ is an inner potential defined solely in a molecular context
(DDL theory) whereas c/J in Eq. 3.20 is defined in a macroscopic context
(thermodynamics). A point of contact between the two can be determined
legitimately by postulating, for example, that the Galvani potential difference across an interface is equal to 0/6 - 0/0, where 0/6 is the PoissonBoltzmann inner potential evaluated at the interface and % is that
evaluated at a point in the aqueous solution phase far from the interface. 14
The concept of model-dependent inner potentials is discussed at length in
Chap. 5.

94

THE SURFACE CHEMISTRY OF SOILS

3.4. ELECTROKINETIC PHENOMENA

The ~_of ~!!,,~e.l~E!!:~~~~..~S.2E.~:li9uid in~!faS~_.!2"_~he~r.i!l,g,.1t~~~

..a.1?p.E~~Ug"J)!j!l.~h!~~tl!l...~~.21!!~~Ou~lillE!J2h~.i1"ler~,e~t~!1..!!~!r.
l~,.it!:~~J!.~!!!;,~;,!'}!'r:.. The fundamental physical assumptions on which the

molecular interpretation of electrokinetic phenomena is based are that an

~!~~~~~~~~~~~~~2*~'m~: i;,jf~~~t~w~~f!A~~f~~~~

p~~ar~t~lc"l"e~su~r'"'1'*'ac"'e"",~t1!h~eriqJrid ph';;e is assumed to be at rest relative to the solid

particle; beyond the plane out into the liquid phase, the liquid moves
relative to the solid particle because of the shearing stress it experiences.
This relative motion perturbs the interfacial region in a manner that one
assumes can be described through the simultaneous application of the
Poisson equation in classical electrostatics and the Navier-Stokes equation
in fluid mechanics.l" Alternatively, one can formulate a description of
electrokinetic phenomena as an application of methods in the thermodynamics of irreversible processes, with no appeal made to specific models of
the interfacial region."? With this approach, of course, detailed information about the molecular properties of an electrified interface cannot be
obtained from an analysis of electrokinetic data.
For soil clay particles, it is often the case that the radius of curvature of
any patch on the particle surface is very much larger than the mean thickness of the interfacial region extending into the soil solution. A perfectly
flat clay particle surface meets this condition exactly, for example, because its radius of curvature is infinite by definition. In this case, the general
description of electrokinetic phenomena in terms of electrostatics and fluid
mechanics is made simpler because there is no perturbation of the interfacial
region except along the direction normal to the particle surface and no distortion of the ion swarm in the liquid phase except that produced by the
charge on the particle before the plane of shear came into being.J" With

Figure 3.3. Geometric aspects of electrokinetic phenomena.


y

SOLID
SURFACE

THE ELECTRIFIED INTERFACE IN SOILS

95

these two physical conditions in mind, one can define an inner potential in
the mobile liquid phase near the solid particle through the Poisson
equation:
d ( soD dljJ)
dx
dx = - p(x)

(x

:2:

d)

(3.26)

where ljJ(x) is the inner potential, p(x) is the volumetric charge density
(coulombs per square meter), and the other symbols are as defined in
connection with Eq. 1.11. As indicated in Fig. 3.3, the coordinate x is
measured from the solid particle surface out into the liquid phase. The
inner potential, ljJ, is subject to the constraints
ljJ(b)

=0

(3.27)

and

where b ~ d is a point in the mobile liquid phase that is far from the plane
of shear located at x = d. Equations 3.26 and 3.27 must be regarded as
model definitions of the inner potential since ljJ(x) cannot be measured by
any model-independent technique.
The other dependent variable of interest is the liquid velocity, V,(x) ,
defined to be a solution of a linearized form of the Navier-Stokes equation:
dP
d ( dVl)
f(x) - dz + dx T/ dx

=0

(x

> d)

(3.28)

where P is the pressure applied to the liquid, f(x) is the external force per
unit volume applied to it, and T/ is its coefficient of viscosity. Equation 3.28
is a mathematical expression of the balance of forces on a differential
element of the liquid under steady-state conditions. The liquid velocity is
subject to the constraints
(3.29a)
and

dV')
( -dx x=b =0

(3.29b)

where U is the velocity in the mobile liquid phase, measured relative to the
particle surface, at a point x = b far from the plane of shear.
The net electric current I, which under steady-state conditions, is
produced by the convection of charged species in the mobile region of the
liquid phase, can be expressed mathematically with the equation
I =

JJ: p(x)vl(x)dxdy .

(3.30)

where frlJl is just the net electric current density through an element of
cross-sectional area dxdy. The mean-value theorem of integral calculus and

THE SURFACE CHEMISTRY OF SOILS

96

Eq. 3.29 can be used to simplify Eq. 3.30:


1= vl(d)

=u

If:

Ir

p(x)dxdy

VI(b)

JJ:

p(x)dxdy

(3.31)

p(x)dxdy

where x = a lies somewhere between the plane of shear and x = b. Now


Eq. 3.30 can also be transformed with the help of Eqs. 3.26 to 3.29 and the
assumption that the mobile liquid phase is a homogeneous, viscous
dielectric medium:

The first step in Eq. 3.32 is the result of substituting Eq. 3.26 into
Eq. 3.30; the second step is an integration by parts; the third step makes
use of Eqs. 3.27 and 3.29a and invokes the assumption that the mobile
liquid phase is a homogeneous dielectric medium; the fourth step is
another integration by parts; the fifth step is the result of Eqs. 3.27 and
3.28 along with the assumption that the mobile liquid phase has a uniform
viscosity coefficient; the sixth step involves the identity
d(dljl)
dx

== dZIjIz dx
dx

and the use of Eq. 3.27; and the last step requires Eq. 3.28 again. !he
'&lY~.QfJl:teJ!1.I.!.~!..E9telltiaJ
. at.1Q~ pl~I!~ gf shear i~ denoted by , ill.Eq, .~, ~~
and is called the zeta. potential of the interface.' . . .To the extent that the liquid phase retains bulk dielectric characteristics
outside the region enclosed by the plane of shear. Eqs. 3.31 and 3.32 lead

THE ELECTRIFIED INTERFACE IN SOILS

97

to a general expression for electrokinetic phenomena:


U

II:

p(x)dxdy =

_eo~(

II: [1 -

o/i)] [f(X) -

~~]dXdY

(3.33)

The mean value theorem can be applied to the right side of this equation in
the same manner as in Eq. 3.31 with the quantity [1 - (o/(x)/()] taken
outside the integral sign. The quantity equals unity when x = b and zero
when x = d. Therefore, Eq. 3.33 reduces to
U

[b
f Ja p(x)dxdy =

e D(

-~

ffba' [f(x)

dPJ
- dz dxdy

(3.34)

where x = a' lies between x = d and x = b. In applications of Eq. 3.34, it


is assumed that the interfacial region is thin enough that VI i U and 0/ ~ 0
so rapidly that the distinction between a, a', and d can be ignored. Under
this assumption, the lower limits of both of the explicit integrals in Eq. 3.34
can be set equal to d. The expression then takes a convenient form that can
be used to describe the four principal electrokinetic phenomena that are
investigated in soil clay suspensions.

where UE is the steady-state velocity of a charged particle moving in


response to an electric field of magnitude E. ~-l.lnits..Ql!:(,1!r,~....m_2s:~y~:,

;;o~~~t:L1g~;~~iH~elrU;~s~;:~li~;~~~!*~~~~~~i~a~~~:T~)~

positive surface charge density at the plane of shear, and negative if UE is


anti parallel with E, indicating a negative surface charge density at the
plane of shear.
In the case of electrophoresis, the only force on the charged particle is
the electric force:
dP
-=0
f(x) = p(x)E
dz
and Eq. 3.34 reduces to
U

[b

Jd

p(x)dxdy =

e D(

I[b

-~ E Jd p(x)dxdy

or, after cancelling the integral from both sides,

U = 6 oD(E
T1

(3.36)

THE SURFACE CHEMISTRY OF SOilS

98

Finally, note that U = - U E because U is the velocity of the liquid phase


relative to the particle and that Eqs. 3.35 and 3.36 then lead to the result
eoD(
u = -

(3.37)

TJ

The sign of u is determined by that of the zeta potential, which in turn


depends on that of the surface charge density at the plane of shear.
Equation 3.37 depends in an essential way on the stipulation that l/J(x) (and
() are solutions of the Poisson equation (Eq. 3.26). It does not require that
l/J(x) be a solution of the Poisson-Boltzmann equation. If that condition is
also imposed, then standard DDL theory may be apj21ied to relate the

surface charge densi~!.~~I::J'l~_~.2~~~ar to (: 19- .. -

--_._:;~~ {ZEoDRT+ c,[eXP(~~~/R1) -

!Jr

(3.38)

where
sgn()

+1

= { -1

if
if

c> 0
0

and the other symbols (except have the same meaning as in Eq. 1.11. If
it is further assumed that the plane of shear is located at the outer
periphery of the outermost surface complexes on the soil particle, then
a, = -aD and Eq. 3.38 becomes:
aD

/
f

= -sgn() {2e oDRT ~cj[exp(-ZjFURT) - 1]

(3.39)

The sum in Eq. 3.39 is over all charged species (with valence ZJ in the
mobile liquid phase. Equations 3.37 and 3.39 are the principal electrophoretic expressions applied to soil clay particles. Clearly, Eq. 3.37 is
dependent on fewer assumptions concerning the structure of the interfacial
region near a soil particle than is Eq. 3.39. However, the present
consensus/" is that the use of DDL theory to derive Eq. 3.39 is a valid step
for 1(' < 0.1 V and electrolyte concentrations below about 10 molom- 3
A sufficient theoretical basis for die use of electrophoresis to measure
the PZC, as discussed in Sec. 3.2, can be developed with Eq. 3.37 and the
single assumption that the plane of shear coincides with the periphery of
the surface complexes on a soil particle. Under this assumption, the
vanishing of aD at the PZC (Table 3.1) implies that the surface charge
density on the plane of shear vanishes as well. This condition and its
consequence, p(x) = 0, then must also obtain on any plane beyond the
plane of shear out into the mobile liquid phase, but Eqs. 1.13 and 3.26
applied to these planes lead to the conclusion that the inner potential,
l/J(x), is equal to a constant everywhere in the mobile liquid phase. This
constant may be set equal to zero, from which it follows that, = 0 and that
u in Eq. 3.37 vanishes at the PZC, as illustrated in Fig. 3.1. Thus it is not

THE ELECTRIFIED INTERFACE IN SOILS

99

necessary to use DDL theory to establish a general relation between the


electrophoretic mobility and the PZC. However, an unverifiable assumption about position of the electrokinetic plane of shear must be made.
This assumption of congruence between the plane of shear and the
periphery of a soil particle has also played a critical role in the quantitative
assessment of the degree of surface complexation by montmorillonite.j"
Table 3.5 lists representative measured values of the electrophoretic
mobility of Na-montmorillonite particles suspended in 10- 4 M NaCI near
pH 6. The compiled data, which show variation over a factor of 2, may be
compared with u = -2.5 0.5 x 10- 8 m2s- 1V-I, the value observed
in Na-montmorillonite suspensions free of background electrolyte.F
Evidently, differences in clay mineral preparation and experimental procedure account for this variability. Regardless of which value of u is
accepted in the range of available data, the use of Eq. 3.39 to interpret the
electrophoretic mobility yields an estimate of 0'0 that lies between 1 and 2
per cent of the absolute value of (Tin for Na-montmorillonite.P Therefore,
if the plane of shear is indeed at the periphery of the Na-montmorillonite
particle and DDL theory applies to the electrolyte in the mobile liquid
phase, it can be concluded from available electrophoretic mobility measurements that about 98 per cent of the Na + cations on the clay particle are
bound into surface complexes. This conclusion is in significant contradiction with the estimate of 23 per cent made in Sec. 1.4 from the results
of negative adsorption experiments on 0- anions in Na-montmorillonite
suspensions. Since Na-montmorillonite suspensions in 10- 4 M NaCI do
not coagulate (Sec. 6.2), the clay particles studied in the electrophoresis
experiments should not have been quasicrystals or other kinds of aggregate
unit. Therefore, it appears that the high estimate of degree of surface
complexation is the result of assuming that the plane of shear and the
periphery of the clay particle coincide. If the plane of shear in fact encloses
a number of layers of immobile water beyond the periphery of the clay
particle, then (TD is underestimated by Eq. 3.39 and the discrepancy can be
explained.
It is very important to note here that DDL theory can give no
information about the location of the clay particle surface because it cannot be applied unambiguously to the region enclosed by the the plane of
shear. For example, if it is assumed that any surface complexes on
Na-montmorillonite occupy only the surface plane in DDL theory (i.e.,
Table 3.5. Some measured values of u for Na-montmorillonite
suspended in 10- 4 M NaCl at pH values near 621

Source

Measurements

Ravina and Zaslavasky


Callaghan and Ottewill

12
30

Low

34

-2.2 0.3
-3.1 0.2
-4.4 0.5

THE SURFACE CHEMISTRY OF SOilS

100

x = 0 in relation to Fig. 3.3), then it is a straightforward matter to show


that, at 298 K in a 0.1 mol m -3 NaCl solution.v'

Iu,1

0~04

(3.40)

where 5, is the distance, in nanometers, between the particle surface and


is in coulombs per square meter. If
the plane of shear and
3
= 0.02Uin = -2 X 10- Cm- 2 for Na-montmorillonite (Eq. 1.22),
then it follows from Eq. 3.40 that 5, = 20 nm. Thus DDL theory predicts
that the plane of shear is more than sixty molecular diameters away from
the plane in which up resides. On the other hand, if it is assumed that
= -UD, then DDL theory cannot be applied consistently in the region
o < x <: d in Fig. 3.3 and Eq. 3.40 is invalidated.
Equation 3.37 can be applied to surface complexes on Ca-montmorillonite without making an assumption about the location of the plane
of shear. 24 The data graphed in Fig. 3.4 indicate that the absolute value of
the electrophoretic mobility of montmorillonite particles suspended in
distilled water decreases as the clay is converted from the sodium form to
the calcium form. This decrease does not commence until about 65 per cent
of the intrinsic surface charge on the clay is balanced by Ca 2 + cations;
thereafter, the decrease is quite rapid. When the charge fraction of Ca 2 +
on the clay is near 0.7, it is known that outer-sphere surface complex
formation between Ca 2 + and the siloxane ditrigonal cavity (Fig. 1.8)
produces quasicrystals containing several crystallites of unit-cell thickness
stacked along the direction of the crystallographic c axis. The Na + cations
that remain on the clay are relegated to the external surfaces of the
quasicrystals, and it is these external surfaces whose charge densities
determine the electrophoretic mobility. Figure 3.4 indicates that, as Na + is

u,

u,

u,

Figure 3.4. The electrophoretic mobility of montmorillonite as a function of the


charge fraction of exchangeable calcium on the clay. 22

WYOMING BENTONITE
0.1% (w/v) SUSPENSION
BAR-ON ET AL. (1970)

>
I

-1.0

<n

(\J

CXl

-2.0

- 3.0

I
o

I...-_ _

0.2~

"-_-"""---~

O.~O

0.7~

1.0

CHARGE FRACTION OF CALCIUM ON CLAY

THE ELECTRIFIED INTERFACE IN SOILS

101

replaced by Ca 2 + on the external surfaces of the quasicrystals, the absolute


value of u decreases rapidly. This decrease is consistent with a decrease in
induced by enhanced surface complexation of Ca 2 + relative to Na + on
the clay. The fact that the value of u for Na-montmorillonite is essentially
unaffected by an increase in NaCI concentration from zero to 0.1 mol.m"?
whereas the value of lui for Ca-montmorillonite drops by 30 per cent over
the same increase in concentration of charge" offers further, indirect
experimental support for this interpretation in terms of surface complexation. Additional indirect evidence that the extent of surface complexation
on Na-montmorillonite is less than on Ca-montmorillonite is provided by
the observations of an increase in the electrophoretic mobility of the
Na-clay after cation exchange with proteins and of no change in the
mobility of the Ca-clay under the same circumstances.P Evidently, exchangeable proteins can compete effectively with adsorbed Na + to form
complexes that screen the siloxane surface charge well and thereby
increase a,. These proteins are less able to displace adsorbed Ca2+, and
their surface complexes produce values of that are the same as those of
the pure Ca-montmorillonite.

a,

a,

Electro-osmosis is the response of an electrolyte solution near an electrified interface to an applied


constant electrk..fid<1. 16.t B In electro-osmosis, only the liquid phase is free
to move (e.g., through a capillary tube or the interstitial space in a plug of
soil clay). The force on the electrolyte ions is the electric force,
f(x) = p(x)E, and Eq. 3.34 reduces to
ELECTRO-OSMOSIS AND THE STREAMING POTENTIAL.

UE O

fI

p(x)dxdy

-eoD(E fIb
T/

p(x)dxdy

or, with cancellation of the integral factor,


UEO

-eoD(E

= _.::.-.::.-

(3.41)

T/

where U == UE O is the electro-osmotic velocity of the mobile liquid phase.


Equation 3.41 predicts that the mobile portion of an electrolyte solution
near a stationary electrified interface moves either parallel or antiparallel
with the direction of an applied electric field E, depending on the sign of
the zeta potential. If the surface charge on the electrokinetic plane of shear
is negative, ( < 0 and the electrolyte solution moves in the same direction
as E. Mechanistically speaking, UE O is parallel with E in this case because
there is a net excess of positive ionic charge in the mobile liquid phase.
After the electric field is applied and steady-state motion commences, the
excess cations and the liquid they drag along with them must be pushed in
the same direction as the applied field. 18 If the surface charge is positive,
( > 0 and UE O is antiparallel with E because a net excess of negative ionic
charge exists in the mobile liquid phase. Although in either case UE O refers
strictly to a point in the mobile liquid phase far removed from the plane of

a,

102

THE SURFACE CHEMISTRY OF SOilS

shear, the assumption of a rapid approach of Vl(X) to its limiting value,


used in the derivation of Eq. 3.34, suggests that almost all of the mobile
liquid phase has the velocity UE O under steady-state conditions. For this
reason, it is common to equate UE O with the volumetric liquid flux density
(volume of liquid transported through unit cross-sectional area per unit
time) measured in response to an applied electric field. 16
The streaming potential is an electric potential difference induced by the
response of an electrolyte solution near an electrified interface to an
~pplied uniform pressure gradient. 16,18 As in electro-osmosis, only the
liquid phase moves in response to the pressure gradient. The convection
current, I, established through this liquid motion can be deduced from
Eq. 3.32 in the special case where [(x) = 0, dP/dz = constant:
I

AeoD( dP

Tf

(3.42a)

dz

where A is the cross-sectional area of the mobile liquid phase perpendicular to the direction of the current. The streaming potential gradient
induced by I when it is not shunted is defined by the equation (without the
usual minus sign because the liquid phase is moving)
I

= AK dVst

(3.42b)

dz

where K is the conductivity of the medium (bulk liquid phase and


interfacial region) through which I flows. Equations 3.42 lead to an
expression for the streaming potential, .i V st :
eoD

(3.43)

.iVst = KTf (.iP

which states that an applied pressure difference .iP induces an electric


potential difference .iV st whose sign depends on that of the zeta potential.
!? terms of mechanism, this pQtj:mtia) djfference arises because the,..ljg,Wd
motion caused by the pressure difference separates the excess charge in the
!!!..C?bile liquid phase from the surface charge it balances on the plane of
shear. If this surface charge is negative,
and a positive potential
gradient is induced by a negative pressure gradient. Stated another way,
the electric field corresponding to the streaming potential gradient,
Est = -dVst/dz, is parallel with the pressure gradient if (is negative and
antiparallel if ( is positive.
There is an essential physical symmetry between electro-osmosis and the
streaming potential in that, for both phenomena, an applied force causes
the mobile portion of a liquid phase near a stationary, electrified interface
to flow. In electro-osmosis, an applied electric force causes the mobile
liquid phase to move with the velocity UE O ' given in Eq. 3.41. The
corresponding volume flow is 0 - VEoA, where A is as defined in
Eq, 3.42a. After multiplying both sides of Eq. 3.41 by A and comparing

THE ELECTRIFIED INTERFACE IN SOILS

103

the result with Eq. 3.42, one can derive the symmetric relationship

r;
Q
_.::..:..--=dl'[tiz

(3.44)

where 1st == - AKdVstldz is the streaming current. Equation 3.44 is the


result of applying the Poisson equation and the Navier-Stokes equation to
an aqueous electrolyte solution near an electrified interface. However, the
same expression can be derived without these model equations through an
application of the Onsager reciprocal relation in the context of the
thermodynamics of irreversible processes.l? Thus, Eq. 3.44 describes a
fundamental physical characteristic of electrokinetic phenomena: the
reciprocity of the ratio of the response to the driving force .
.In soils, precise measurements of electro-osmotic flows and streaming
potentials are difficult to make, but the available data/" suggest that
IVEolEI = 10- 8 m2s- IV-1, which agrees wi
ical values of the electrophoretic mobilities of soil clay particles, and t
1.11 Vstll aP is of the
7
order of a few millivolts per bar (= 10- V, Pa -1). Sin e a typical value of
K is Eq. 3.43 is around 0.1 C'm- 1s- 1V- 1
r moist soils,26
1
1V- 1
7
2N1sKlaVstl/aP = 10- C'mx 10- V'm
= 10- 8 m2s- 1V- 1
(lC = Nvm- V-I). This approximate result shows that Eq. 3.44 is satisfied
in the form
KiaVstl
IVEol
aP E
(3.45)
which can be deduced from Eqs. 3.41 and 3.43.
Electro-osmosis and the streaming potential play an important role in
the response of a soil clay layer saturated with an electrolyte solution to an
imposed gradient in the concentration of the electrolyte." Since an
electrified clay-aqueous solution interface is present, both the concentration and the mobility of the electrolyte cation in the clay material differ
from those of the electrolyte anion. When the concentration gradient is
imposed, these differences between the two kinds of ion produce different
ion fluxes through the clay layer. The different fluxes in turn contribute to
an electric potential difference across the clay.28 One effect of the induced
electric potential difference is electro-osmotic flow. However, if no liquid
flow across the clay layer is permitted, there develops a pressure gradient
that opposes electro-osmosis. This induced pressure gradient contributes
to a streaming current through the clay. If no electric current through the
clay layer is permitted either, then the applied concentration gradient
determines the induced gradients of electric potential and pressure
uniquely as the solutions of the simultaneous linear equations/"
dP
dz

o=

o=

LEY -

dP
dz

dC
L YD dz

+ LED

dC
dz

dV
LYE dz

+ LE

dV
dz

104

THE SURFACE CHEMISTRY OF SOILS

where the L coefficients are constant parameters that characterize the soil
clay-electrolyte solution mixture and C is the electrolyte concentration.
The first equation describes the flow of water through the clay layer, and
the second describes the electric current through the clay. The coefficient
LYE is equivalent to the coefficient of -E in Eq. 3.41; the coefficient LEY is
equivalent to the coefficient of A df/dz in Eq. 3.42a. The equality
LYE = LYE is another way of expressing the reciprocity in Eq. 3.45.
When a soil particle bearing an electrified interface settles in an aqueous solution under the force of gravity, a
plane of shear develops around the particle just as it does in electrophoresis. As the particle settles, the portion of the interfacial region
enclosed by the plane of shear moves with it but the remainder is left
behind, and this separation of interfacial charge gives rise to an e ctric
potential difference called the sedimentation potential. For a suspension 0
soil particles that do dot interact with one another (i.e., that settle
independently), the force per unit volume acting on the particles is
THE SEDIMENTATION POTENTIAL.

[(x) = -cPdpg

(3.46)

where cP is the volume fraction of particles in the suspension, dp is the


difference in mass density between a particle and the surrounding liquid
phase, and g is the gravitational acceleration. Equation 3.46 may be
substituted into Eq. 3.32 (with dP/dz = 0 and neglect of /I(x)/,) to derive
an expression for the convection current produced by sedimentation:

1=

-soD

,AcPdpg

(3.47)

T/

A definition exactly analogous to Eq. 3.42b can be given for the gradient of
the sedimentation potential, with the result that Eq. 3.47 is transformed to
the expression
(3.48)
(In this case, a minus sign is included in the relation between I and
dVsed / dz because it is the liquid phase that does not move. Therefore, no
minus sign appears in Eq. 3.48.) Equation 3.48 predicts that the gradient
of sedimentation potential has the same sign as , and is proportional to cPo
These predictions have been verified experimentally for a variety of
colloids.i" but apparently no study has been done on soil clay particles. It
may be noted in passing that Eq. 3.48 applies equally to particles undergoing centrifugation if g is replaced by the centrifugal acceleration. Since
centrifugal accelerations acheived in laboratory studies of ion adsorption
phenomena on soil clays often are tens of thousands of times larger than g,
the gradients of Vied generated can be very large. It is possible that these
potential gradients and the concomitant shearing away of part of the ion

THE ELECTRIFIED INTERFACE IN SOILS

105

swarm can significantly alter the ion distribution in the interfacial region. If
this disruption occurs, centrifugation may introduce an artifact into the
measurement of ion adsorption.
Several general aspects of electrokinetic phenomena, as summarized in
Eqs. 3.36, 3.41, 3.43, and 3.48, should be emphasized at this juncture.
First, since ( is an inner potential, it cannot be measured directly. It can
only be inferred from electrokinetic data through model-dependent equations. The fact that values of ( calculated from the results of different
electrokinetic experiments on the same interface are in agreemenr'" does
not demonstrate an objective existence of the zeta potential because all the
experiments are described by the same general equation (Eq. 3.34).
Second, electrokinetic data can be given molecular significance through the
zeta potential simply by assuming the existence of a plane of shear and the
applicability of the Poisson and Navier-Stokes equations. Diffuse double
layer theory is not required. Therefore, electrokinetic data offer no direct
experimental support for the DDL assumptions. Even the assumption that
a definite plane of shear exists can be modified. Consider, for example,
electro-osmosis. This phenomenon is the result of an applied electric force
balanced by the viscous force under steady-state conditions (Eq. 3.28):
p(x)E

d (T/ dVl)
+ dx
dx = 0

(3.49a)

With the help of Eq. 3.26 this expression takes the form

E~(eoD:) = ~ (T/~l)

(3.49b)

A single integration of both sides of this equation and the evaluation of the
constant of integration with the help of Eqs. 3.27 and 3.29b give the
differential equation

dl/J
EeoD/
dx
This expression can be divided by

de,
T/dx

T/ on both sides and then

UEo = - eoE

(3.49c)
integrated:

{D

I T/
o

- dl/J

(3.49d)

where the boundary conditions in Eqs. 3.27 and 3.29 (with U = UE O ) have
been noted again. Equation 3.49d is a generalization of Eq. 3.41 to permit
D and T/ to be functions of the potential l/J (or of the coordinate x). The
observed electro-osmotic velocity, UE O , thus depends on the variability of
both the dielectric and the viscosity properties of the liquid phase. The
possibility that UE O results only from continuously variable D and 71,
without the existence of a plane of shear, is consistent with Eq, 3.49d':~o

----

106

THE SURFACE CHEMISTRY OF SOilS

This conclusion, which can be extended to the other electrokinetic


phenomena considered in this section, illustrates the tenuous nature of any
assumption concerning the exact location of the plane of shear. Any
position of the plane at x = d in Fig. 3.3 is, in principle, consistent with the
results of an electrokinetic experiment.
Finally, it should be evident that the principal utility of electrokinetic
data is qualitative and comparative. The demonstration that an electrokinetic phenomenon exists is a demonstration that an electrified interface ~
exists. This broad implication is conclusive. In a quantitative sense,
however, measurements of electrokinetic phenomena on the same kind of
interface under different conditions have only a comparative value.
Changes in the zeta potential as the composition of the interfacial region is
varied through adsorption can be a useful guide to molecular interpretation
even though' itself has little objective significance.
3.5. NEGATIVE ADSORPTION

The general mechanism of negative ion adsorption is described in Sec. 1.4,


where the relevance of this phenomenon to the measurement of specific
surface area is discussed. The prototypical experiment for observing negative adsorption involves the effect of a suspended soil clay on the concentration of an ion whose valence has the same sign as that of the
charge on the soil clay particles, O"p (Eq. 3.2). If O"p is not zero, the
concentration of the co-ion is larger in a supernatant solution in contact
with an equal volume of suspension of the solid than it was before the solid
was suspended. This strictly macroscopic phenomenon can be characterized quantitatively through the definition of the relative surface excess of an
ion i in a suspension:
(3.50)

where n, is total moles of ion i in the suspension per unit mass of solid, Mw
is total kilograms of water in the suspension per unit mass of solid, m, is the
molality of ion i in the supernatant solution, and S is the specific surface
area of the suspended solid. Thus rf w ) is the excess moles of the ion (per
unit area of suspended solid) relative to an aqueous solution containing M w
kilograms of water and the ion at molality m.. The chemical foundation of
the definition of the relative surface excess is examined in Sec. 4.1. For the
present discussion, it is sufficient to note that r}w) can be positive, zero, or
negative, in principle, and that the condition
(3.51)

is a formal, macroscopic definition of negative adsorption for any ion i. As


a numerical example, suppose that a soil clay with a specific surface area of
2 x 104 m2kg- 1 is suspended in a solution of NaCi following the prototypical two-chamber experiment described in Sec. 1.4. The chamber

THE ELECTRIFIED INTERFACE IN SOILS

107

containing the soil clay suspension is found to hold 0.6 kilogram of water
and 2.8 millimoles of CI per kilogram of clay. The supernatant solution in
contact with the suspension through a membrane permeable to water and
ions is 0.007 molal in chloride. Therefore, according to Eq. 3.50,
(w) _
-

r Cl

0.0028 - (0.6 x 0.007)


2

104

-2

mol-rn

= - 7 X 10- 8 mol-m F

and the chloride ion is said to be negatively adsorbed by the soil clay.
It must be emphasized that the definition of negative adsorption
epitomized by Eqs. 3.50 and 3.51 is strictly macroscopic and does not
depend in any way on the concepts of DDL theory applied in Sec. 1.4. If a
DDL theory interpretation of Eq. 3.50 is desired, it can be developed
through the definitions"
n, == ( ci(x)d 3x/m s

Jsu

u; ==

Pw (

Jsu

d 3x / m s

(3.52a)
(3.53b)

where Ci(X) is the concentration (moles per unit volume) of ion i in the
aqueous solution portion of a suspension containing m; kilograms of soil
clay, Pw is the mass density of water in the suspension, and the integrals
extend over the entire suspension volume. Under the assumption that
Pwmi = Ci'" (the concentration of ion i in the supernatant solution),
Eqs. 3.50 and 3.52 can be combined to produce

r~w) =

Sl ( [c;(x) - ci",]d 3x
m; Jsu

(3.53)

Equation 3.53 provides a molecular-level interpretation of Eq. 3.50. The


exclusion volume, introduced in Eq. 1.9, is related to Eq. 3.53 through the
macroscopic definition
=
Vex-

sr(w)/c
i
ioo

(3.54)

Equation 3.54 can be interpreted on the molecular level with the help of
Eq.3.53:
(3.55)
In conventional DDL theory applied to an electrolyte solution bathing
solid particles with planar surfaces, Eq. 3.55 simplifies to
VC K = SE

L'" [1 - c~~:)] dx

(3.56)

THE SURFACE CHEMISTRY OF SOilS

108

where x = 8 defines the plane where the ion swarm is in contact with solid
particle and SE is the surface area of this plane divided by m.. Equation 3.56 can be developed further in DDL theory as indicated following Eq. 1.10. None of this development is necessary to the experimental
description of negative adsorption, of course. That description depends
only on Eqs. 3.50 and 3.51.
As with electrokinetic phenomena, the existence of negative adsorption
implies the existence of an electrified interface. The behavior of this
interface toward charged particles can always be investigated with the help
of a particular molecular model, such as DDL theory, but it is useful to see
how much information can be obtained without a detailed model, in
keeping wih the spirit of the previous sections in this chapter. Consider, for
example, the application of thermodynamics to the prototypical twochamber.experiment on negative adsorption. If the very small osmotic
pressure created by the suspended soil clay is neglected, the activity of any
electrolyte in the two chambers is the same in both the suspension and the
supernatant solution: 14
(MaLb)su = (MaLb)so

(3.57)

where MaLb(aq) is the electrolyte (M = metal; L = ligand) and the su and


so have the same meanings as in Eq. 3.18b. Both of the electrolyte
activities in Eq. 3.57 can be measured electrochemically without making
unverifiable extrathermodynamic assumptions.I' Thus both activities are
well-defined macroscopic quantities. They can be partitioned further with
the use of mean ionic activity coefficients: 14
(3.58)
where y is a mean ionic activity coefficient and
m

==

(mM mt)1/(a+b)

(3.59)

and mi. being the total molalities of the metal M and ligand L.
Equation 3.58 is an exact result in thermodynamics, but it cannot be used
to characterize negative adsorption without making some kind of assumption as to the nature of m";:.
Suppose that M = Na, a = 1, L = CI, and b = 1 in Eq. 3.57. Suppose
further that both Na + and Cl" are dissociated fully from the soil clay
particles in the suspension and that the only important contribution to up is
a negative uo, the surface density of structural charge. These conditions
apply reasonably well to a suspension of montmorillonite clay particles in a
dilute solution of NaCl at pH 7.0.32 In this case, it is possible to write the
following relationship between m~a and mel:
mM

su
mNa

su
mCi -

uoSm s
FM

36Oa)

where -uoS/ F represents the cation exchange capacity of the soil clay
particles. In the suspension, it is possible that the contributions of Na of and

THE ELECTRIFIED INTERFACE IN SOILS

Cl" to

'Y~u

109

are compensating in such a way that


(3.60b)

Equation 3.60b is a model assumption that can be verified experimentally


given Eqs. 3.58, 3.59, and 3.60a. For the present discussion, Eq. 3.60 will
be taken to constitute what may be termed the Babcock model of the
suspension. 33
The equilibrium condition Eq. 3.58 applied to the Babcock model yields
the expression
mSu
mSu
Cl ( Cl

uosms) = m 2
FMw

(3.61)

where m = m~ is the mean ionic molality of NaCl in the supernatant


solution. The negative adsorption of Cl" by the soil clay is given by
Eq.3.50:

r CI(w)

_
-

Mw(m~

- m)

---''-'-'-:':'''_--'-

Sm;

(3.62)

For convenience, the definitions


(w )
11m == S msr Cl

Q==

Mw

-uoSm s
FMw

(3.63)

can be introduced into Eq. 3.61 to give an equation for 11m:

(11m + m)(l1m + m + Q) = m 2
The solution of this quadratic equation is

.::1m

-em + tQ) + ctQ 2 + m 2)1 / 2

(3.64)

where the positive square root is chosen because 11m must vanish with both
m and Q. Eq. 3.64 shows that, as m increases, 11m decreases from 0 to
the asymptotic value -tQ = uoSmsl2FMw . The asymptotic result can be
derived by writing
(tQ2 + m 2)1 /2 = m[1 + (Q/2mf]1 / 2
in Eq. 3.64 and noting that this term approaches m as m becomes
arbitrarily large. Therefore,
lim
mtoo

rCl(w) =

Mw lim 11m =
Smsmt oo

(3.65)

according to the Babcock model.


The chemical significance of Eq. 3.65 can be understood after Eq. 350 is
applied to Na" to derive the expression
r(w)

Na

= Mw

(m~a - m)
Sm

(3.66)

THE SURFACE CHEMISTRY OF SOilS

110

It follows from Eqs. 3.60a, 3.62, and 3.66 that


r(w) _
Na

r(w)
CI

-(T
= __
0

(3.67)

in the Babcock model. The (positive) adsorption of Na + minus the


(negative) adsorption of CI- must always equal minus the surface charge
density of the soil clay expressed in moles per square meter. Equation 3.65
shows that, according to the model, the lowest possible value of r~'r) is half
the surface charge density of the soil clay. Therefore, by Eq. 3.67, the
corresponding smallest possible value of r~l is minus half the surface
charge density of the clay. Under these conditions, half the cation
exchange of the clay is satisfied by the positive adsorption of Na + and half
is satisfied by the negative adsorption of CI-. As the electrolyte concentration decreases, less and less of the surface charge is neutralized by a deficit
of chloride anions.

NOTES
1. G. Sposito, The operational definition of the zero point of charge in soils, Soil
Sci. Soc. Am. J. 45: 292 (1981).
2. These kinds of experiments are described in A. M. James, Electrophoresis of
particles in suspension, Surface and Colloid Science 11:121 (1979).
3. Speciation models for aqueous solutions are discussed in Chap. 3 of
G. Sposito, The Thermodynamics of Soil Solutions. Clarendon Press, Oxford,
1981.
4. See, e.g., D. H. Everett, Manual of Symbols and Terminology for Physicochemical Quantities and Units. Appendix II: Definitions, Terminology and
Symbols in Colloid and Surface Chemistry. Butterworths, London, 1972. When
the PZC is measured by an electrokinetic experiment (Sec. 3.4), it is often
termed an isoelectric point (IEP). However, other definitions of the IEP are
used in the soil chemistry literature. 1
5. See, e.g., Chap. 6 in R. J. Hunter, Zeta Potential in Colloid Science.
Academic Press, London, 1981.
6. Methods for measuring the PZSE and the PZNC in soils are discussed in
Chap. 6 of G. Uehara and G. Gillman, The Mineralogy, Chemistry, and
Physics of Tropical Soils with Variable Charge Clays. Westview Press, Boulder,
Colo., 1981.
7. Sources of data: c.-P. Huang and W. Stumm, Specific adsorption of cations
on hydrous y-Al z0 3 , J. Colloid Interface Sci. 43:409 (1973). B. Bar-Yosef,
A. M. Posner, and J. P. Quirk, Zinc adsorption and diffusion in goethite
pastes, J. Soil Sci. 26: 1 (1975). L. C Bell, A. M. Posner, and J. P. Quirk,
The point of zero charge of hydroxyapatite and fluorapatite in aqueous
solutions, J. Colloid Interface Sci. 42: 250 (1973). A. P. Ferris and
W. B. Jepson, The exchange capacities of kaolinite and the preparation of
homoionic clays, J. Colloid Interface Sci. 51:245 (1975).
8. S. S. Wann and G. Uehara, Surface charge manipulation of constant surface
potential soil colloids. I: Relation to sorbed phosphorus, Soil Sci. Soc. Am. J.
42:~6~ (1978).

THE ELECTRIFIED INTERFACE IN SOILS

111

9. This point is discussed in the context of double-layer theory in M.A.F. Pyman,


J. W. Bowden, and A. M. Posner, The movement of titration curves in the
presence of specific adsorption, Aust. J. Soil Res. 17: 191 (1979).
10. E. Tessens and S. Zauyah, Positive permanent charge in Oxisols, Soil Sci. Soc.
Am. J. 46: 1103 (1982). Equation 3.13 was derived in the context of DDL
theory in G. Uehara and G. P. Gillman, Charge characteristics of soils with
variable and permanent charge minerals. I: Theory. II: Experimental, Soil Sci.
Soc. Am. J. 44: 250 (1980).
11. M.A.F. Pyman, J. W. Bowden, and A. M. Posner, The point of zero charge
of amorphous coprecipitates of silica with hydrous aluminum or ferric hydroxide, Clay Minerals 14:87 (1979).
12. The clearest introduction to the electrochemical potential is given by its creator
in Chap. 8 of E. A. Guggenheim, Thermodynamics (North-Holland, Amsterdam, 1967). Applications to soil solutions are reviewed in Chap. 4 of
G. Sposito, op. cit? The issue of electric potentials near interfaces is discussed
in detail in R. Parsons, Equilibrium properties of electrified interphases,
Modern Aspects of Electrochemistry 1: 103 (1954).
13. R. Parsons, Manual of symbols and terminology for physicochemical quantities and units. Appendix III: Electrochemical nomenclature, Pure Appl.
Chem. 37: 500 (1974).
14. For a discussion of these thermodynamic parameters, see Chaps. 1 and 2 in
G. Sposito, op. cit.3
15. The DDL theory is discussed critically in J.T.G. Overbeek, Electrochemistry
of the double layer, in Colloid Science (H. R. Kruyt, ed.), Vol. I. Elsevier,
Amsterdam, 1952.
16. The most thorough modern introduction to the physical basis of electrokinetic
phenomena is given in Chap. 3 of R. J. Hunter, op. cit? See also
J.T.G.Overbeek. 18
17. P. H. Groenevelt and G. H. Bolt, Nonequilibrium thermodynamics of the
soil-water system, J. Hydrol. 7: 358 (1969).
18. These conditions are discussed carefully in J.T.G. Overbeek, Electrokinetic
phenomena, in Colloid Science (H. R. Kruyt, ed.), Vol. I. Elsevier, Amsterdam, 1952.
19. See, e.g., D. C. Grahame, Diffuse double layer theory for electrolytes of
unsymmetrical valence types, J. Chem. Phys. 21: 1054 (1953).
20. G. R. Wiese, R. O. James, D. E. Yates, and T. W. Healy, Electrochemistry
of the colloid-water interface, in Electrochemistry (J. O'M. Bockris, ed.),
Butterworths, London, 1976.
21. I. Ravina and D. Zaslavsky, Nonlinear electrokinetic phenomena. Part II:
Experiments with electrophoresis of clay particles. Soil Sci. 106:94 (1968).
I. C. Callaghan and R. H. Ottewill, Interparticle forces in montmorillonite
gels, Faraday Disc. Chem. Soc. 57: 110 (1974). P. F. Low, The swelling of clay.
III: Dissociation of exchangeable cations, Soil Sci. Soc. Am. J. 45: 1074 (1981).
22. P. Bar-On, I. Shainberg, and I. Michaeli, Electrophoretic mobility of montmorillonite particles saturated with Na/Ca ions, J. Colloid Interface Sci.
33:471 (1970). R. D. Harter and G. Stotzky, X-ray diffraction, electron
microscopy, electrophoretic mobility, and pH of some stable smectite-protein
complexes, Soil Sci. Soc. Am. J. 37: 116 (1973). S. L. Swartzen-Allen and
E. Matijevi~. Colloid and surface properties of clay suspensions. II: Electrophoresisand cation adsorption of montmorillonite. J. Colloid Interface Sci.
so: 143 (llJ75).

112

THE SURFACE CHEMISTRY OF SOilS

23. See, e.g., Chap. 1 in G. H. Bolt, Soil Chemistry. B: Physico-Chemical


Models. Elsevier, Amsterdam, 1979.
24. See P. Bar-On et al., op. cit.,22 and 1. Ravina and D. Zaslavski, op. cit.21
25. R. D. Harter and G. Stotzky, op. cit.22
26. 1. Ravina and D. Zaslavsky, Nonlinear electrokinetic phenomena. I: Review
ofliterature, Soil Sci. 106: 60 (1968). See also Chap. 11 in G. H. Bolt, op. cit.23
27. D. E. Elrick, D. E. Smiles, N. Baumgartner, and P. H. Groenevelt, Coupling
phenomena in saturated homo-ionic montmorillonite. I: Experimental, Soil
Sci. Soc. Am. J. 40: 490 (1976). P. H. Groenevelt and D. E. Elrick, Coupling
phenomena in saturated homo-ionic montmorillonite. II: Theoretical, Soil Sci.
Soc. Am. J. 40: 820 (1976). P. H. Groenevelt D. E. Elrick, and T.J.M. Blom,
Coupling phenomena in saturated homo-ionic montmorillonite. III: Analysis,
Soil Sci. Soc. Am. J. 42: 671 (1978). See also W. D. Kemper, 1. Shainberg, and
J. P. Quirk, Swelling pressures, electric potentials, and ion concentrations:
Their role in hydraulic and osmotic flow through clays, Soil Sci. Soc. Am. J.
26: 229 (1972).
28. Contributions to the observed electric potential difference are also made from
inherent differences in the measuring electrodes and, more important, from the
Nernstian response of the electrodes to the imposed concentration difference.
29. See, e.g., Chap. 3 in J. T. Davies and E. K. Rideal, Interfacial Phenomena.
Academic Press, New York, 1961.
30. The general question of the interpretation of ( is discussed at length in
S. S. Dukhin and B. V. Derjaguin, Electrokinetic phenomena, Surface Colloid Sci. 7: 1 (1974). See especially pp. 29-32 and 53-55.
31. I. Shainberg and A. Caiserman, Electrochemical potential of NaCl in Namontmorillonite suspensions, Soil Sci. 104:410 (1967). See also Chap. 4 in
3
, op. cit.
G . Sposito
32. For other clay minerals, UH contributes importantly to up and the positive
adsorption of anions cannot be neglected. See, e.g., J. P. Quirk, Negative and
positive adsorption of chloride by kaolinite, Nature 188: 253 (1960).
33. The model is described in detail by K. L. Babcock, Some characteristics of a
model Donnan system, Soil Sci. 90: 245 (1960).
FOR FURTHER READING

K. L. Babcock, Theory of the chemical properties of soil colloidal systems at


equilibrium, Hilgardia 34: 417 (1963). Section II of this classic review describes the
Babcock model of negative adsorption in great detail.
G. H. Bolt, Soil Chemistry. B: Physico-Chemical Models. Elsevier, Amsterdam,
1979. Chapter 7 of this comprehensive textbook reviews negative anion adsorption
experiments and their interpretation in DDL theory.
R. J. Hunter, Zeta Potential in Colloid Science. Academic Press, London, 1981.
This excellent monograph gives a thorough, modern introduction to all aspects of
electrokinetic phenomena, both experimental and theoretical.
H. R. Kruyt, Colloid Science. Elsevier, Amsterdam, 1952. Chapters IV and V in
Vol. I, of this classic treatise, written by J.T.G. Overbeek, provide a thorough,
high-level introduction to the properties of the electrified interface discussed in the
present chapter.

,:'

'\

,\

4
INORGANIC AND ORGANIC
SOLUTE ADSORPTION IN SOILS

4.1. THE ADSORPTION ISOTHERM

Adsorption is the process through which a net accumulation of a substance


occurs at the common boundary of two contiguous phases. I The study of
adsorption in soils is characterized by three laboratory operations that
define the net accumulation of a substance at the interface between solid
soil particles and a contiguous fluid: (1) reaction of the soil with a fluid of
prescribed composition for a prescribed period of time, (2) isolation of the
soil from the reactant fluid phase, and (3) chemical analysis of the soil
and/ or the reactant fluid phase. Step_1 ca!1 take place either with the fluig
phase at rest relative to the soil particles ("batch process") or with the fluist
phase in uI.!.iform motion relative to the soil particles ('1!2':.':.-.tl!!:.ou&.1l,
erocess"). The reaction tIme should be long enough to permIt a close,
approach to thermodynamic e uilibrium but short enough to prevent
IOns. tep 2 is usually carried out in batch processes
unwanted side
ihrougll the applicatIon of centrifugal or gravitational force. It is understood that some of the reactant fluid phase is always entrained with the soil
in this kind of separation.f
The quantitative description of adsoption as a purely macrosc,?pic
heno
.
.
hrou h the conce t of relative su ace excess.I
T,his quantity, denoted by the symbol r iI, is the number of excess moles of
'!. su6stance i per unit area of adsorbing soil solids, determined re!~tiye !o a
datum fluid phase that contains the same number of moles Of a reference
substance j as are in the soil. The concentration of i in the datum fluid
phase is the same aSfhat in the reactant fluid phase separated from the soil
at the end of the adsorption experiment. In mathematical terms, rp) is
defined by the equation
.

~.-_----

. 0 .

reac

(4.1)

THE SURFACE CHEMISTRY OF SOilS

114

where n, and nj are the moles of i and j per unit mass of soil solids, Xi and Xj
are the mole fractions of i and j in the reactant fluid phase after it has been
separated from the soil, and S is the specific surface area of the soil solids.
(The mole fraction of a substance in a solution is the ratio of moles of the
substance in the solution to the total number of moles of all substances in
the solution.) This definition of relative surface excess assumes that neither
i nor j enters into the structure of the adsorbing soil solid phase. 1 Note that
r~j) can be either positive or negative because it describes a net accumulation. Also note that there is no adsorption of i when the condition

n, Xi
-=nj
Xj

(4.2)

is fulfilled. The role of reference substance j is evident on the basis of the


fact that
(4.3)
By definition, there is no surface excess ofj. Thus r}j) is a surface excess of
i referred to an interface at which no net accumulation of j occurs."

The natural choice for j in Eq. 4.1 is the substance having the largest
concentration in the fluid phase from which a soil adsorbs the constituent i,
provided there is no significant competition between i and j for adsorbing
surface. If the reactant fluid phase is soil air, the reference substance
should be nitrogen gas; if it is the soil solution, the reference substance
should be liquid water. In the case of a soil adsorbing a vapor from air, it is
expected that the number of moles of nitrogen gas in the interstitial space
of the soil will be. negligible because of the low density of soil air. For
example, a cubic centimeter of soil with a porosity of 0.5 contains only
about 1.5 micromoles of (nonadsorbing) nitrogen gas but could contain 104
times as many moles of adsorbed water vapor. Thus nj in Eq. 4.1 can be
neglected with no loss of accuracy, and the relative surface excess can be
expressed by the equation
I', "'" ni
I

(4.4)

for any substance i adsorbed strongly by the soil particles. If nitrogen gas
itself, or some other weakly adsorbing vapor, is to be adsorbed by the soil
(e.g., to determine its specific surface area), then only that gas should be
present in the soil interstitial space and Eq. 4.4 is still the expression to use
to calculate the amount adsorbed since the number of moles of nonadsorbed gas in the soil will again be negligibly small. 1
In the case of a soil adsorbing a dissolved solute from an aqueous
solution, Eq. 4.1 can be reduced to
(4.5)

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

115

where M w is the mass of water in the soil per unit mass of soil solids (the
gravimetric water content) and m; is the molality of the adsorbing
substance i in the aqueous solution isolated from the soil after reaction.
The product Mwmj can be represented accurately by (JcJ Pb, where (Jis the
volumetric water content of the soil, Pb is its bulk density, and c, is the
molarity of i in the isolated aqueous solution phase. If the adsorption
reaction is initiated by immersing 1 kilogram of a soil into M T w kilograms
of water containing i at molality m?, then the law of conservation of mass
requires that

n, = m? M Tw

mj(M T w

Mw )

This condition can be introduced into Eq. 4.5 to derive the useful
expression
(4.6)
where Sm, == m? - m, is the change in molality caused by adsorption.
Equation 4.6 provides the theoretical basis for the common method of
measuring the amount adsorbed through chemical analysis of the aqueous
solution phase only. 2
Equation 4.5 represents the surface excess of substance i relative to an
aqueous solution that contains M w kilograms of water plus substance i at
molality m.. This surface excess is assigned to a surface at which there is no
net accumulation ofwater. If water in the interstitial space is not adsorbed
(in the sense defined in Chap. 2), then this surface can be taken as
congruent with the geometric boundaries of the adsorbing soil particles. If
some of the interstitial water is adsorbed, say, within the region bounded
by a surface 1.0 nm from the boundary of a soil particle, then the surface of
zero net accumulation of water could differ slightly from the soil particle
surface.
As a numerical example of the application of Eq. 4.5, consider a
montmorillonitic soil with a gravimetric water content of 0.6 kg kg- 1 and
from which 110 mmol of calcium per kilogram of soil has been extracted
after a brief reaction with an aqueous solution whose final calcium molality
was 5 x 10- 3 m. The specific surface area of the soil is 180 m2g- 1 . The
relative surface excess of calcium in the soil is then
w
r ()
Ca

(0.11 mol kg-I) - (0.6 x 0.005) mol kg ?


(180 m
g-kg"")

= -'-- - - --='---,:----=-'- - - '---,---- -=:.2g- 1)(1000

= +0.594 JLmolm- 2

Frequently the value of qj == r}W) S is reported instead of the relative


surface excess. In the present example, qCa = +0.107 mol kg- 1 is the
positive excess moles of calcium per unit mass of soil. An example of the
application of Eq. 4.5 to calculate the negative surface excess of chloride in
a soil clay is given in Sec. 3.5.

116

THE SURFACE CHEMISTRY OF SOilS

It is ~IEp...2E.".pr~cti~~~~~ig!!~J..'Y~~.~ soil aQd an


aqueous solution at a controlled temperature and applied pressure. T~
resulting adsorption data are either (r~w>,mi) or (qi,Ci) eairs that can be
e!!!~d against one another with eith~r r~w) or q i l!L1hsU!~J2~!"~~nt
variabi'e:"This kind of graph is called an adsorption isotherm. Adsorption
isotherms for many substances in soils have been determined over the
past century since the pioneering work of the soil chemist J. M. van
Bemmelen." Enough data for the case of positive adsorption are available
to permit a broad classification of isotherms according to initial slope. 5
Illustrative examples of this classification system are shown in Fig. 4.1.
The S-curve isotherm is characterized by an initial slope that increases
with the concentration of a substance in the soil solution. This property
suggests that the relative affinity of the soil solid phases for the substance at
low concentration is less than the affinity of the soil solution. In the
example of copper adsorption by the Altamont soil given in Fig. 4.1, it is
believed that natural organic compounds in the soil solution form strong,
non adsorbing complexes with the metal. After the complexing capacity of
these compounds is exceeded through an increase in the amount of copper
added to the soil solution, the solid particles in the soil gain in the
competition and begin to adsorb copper ions. Thereafter the isotherm
takes on its characteristic S shape. In some. instances, especially when
organic compounds are being adsorbed, the S-curve isotherm is the result
of cooperative interactions among the adsorbed molecules. These interactions (e.g., surface polymerization or stereochemical interactions) cause
the adsorbate to become stabilized on a solid surface, and they produce an
enhanced affinity of the surface for the adsorbate as its surface excess
increases.
The L-curve isotherm is characterized by an initial slope that does not
increase with the concentration of a substance in the soil solution. This
property is the result of a high relative affinity of the soil solid phases for
the substance at low concentrations coupled with a decreasing amount of
adsorbing surface as the surface excess of the adsorbate incteases. The
example of o-phosphate adsorption in Fig. 4.1 illustrates a universal
L-curve feature: an isotherm that is concave to the concentration axis
because of the combination of affinity and steric factors.
The H-curve isotherm is an extreme version ofthe L-curve isotherm. Its
characteristic large initial slope (in comparison with the L-curve isotherm)
suggests a very high relative affinity of the soil solid phases for an adsorbing
substance. This condition is usually produced either by highly specific
interactions between the solid phase and the adsorbing substance or by
significant van der Waals interactions contributing to the adsorption
process. The example of cadmium adsorption at very low concentrations
by a kaolinitic soil shown in Fig. 4.1 illustrates an H-curve isotherm caused
by very specific interactions. Large organic molecules and inorganic
polymers provide H-curve isotherms resulting from van der Waals interactions.

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOilS


5

117

L -curve

S-curve

01

-'"

"0
E
E

01

-'"

"0
E
E

Anderson sandy
cloy loom
pH 6.2
25C
r = 0.02M.

Q.

0"

Altamont cloy loom


pH 5.1 25C
r = O.OIM

::>

0"

00

CUT (rnrnol
H-curve
0.60

150

C-curve

01

-'"

"0 0.40

E
E

"0

0"

01

Boomer loom
pH 7.0 25C

Il':l0.005M.

0.20

100

-'"

::l..

0"

50

Har-Barqan cloy
parathion adsorption
from hexane
50% RH hydration

0.05

0.10

0.15

0.20

0.25

Cdr (mmol m- 3)

Figure 4.1. General classes of adsorption isotherms. S curve, data courtesy of


C. S. LeVesque; L curve, data from I.C.R. Holford et a/.8 ; H curve, data from
J. Garcia-Mirayagaya and A. L. Page, Sorption of trace quantities of cadmium by
soils with different chemical and mineralogical composition Water, Air, and Soil
Pollution 9: 289 (1978); C curve, data from B. Yaron and S. Saltzman, Influence of
water and temperature on adsorption of parathion by soils Soil Sci. Soc. Am. J.
36:583 (1972).

The C-curve isotherm is characterized by an initial slope that remains


independent of the concentration of a substance in the soil solution until
the maximum possible adsorption. This kind of isotherm can be produced
either by a constant partitioning of a substance between the interfacial
region and an external solution or by a proportional increase in the amount
of adsorbing surface as the surface excess of an absorbate increases. The
example of parathion (diethyl p -nitrophenyl monothiophosphate) adsorption in Fig. 4.1 shows constant partitioning of this compound between
hexane and the layers of water on a soil at 50 per cent relative humidity.
The adsorption of amino acids by Ca-montmorillonite also exhibits a

40

THE SURFACE CHEMISTRY OF SOilS

118

C-curve isotherm, this time because the adsorbate can penetrate the
interlayer regions of quasicrystals, thereby creating new adsorbing surface
for itself. 5
T~ L-curve isotherm is by far the ~~~.ommonlyencountered in
the literature of soil chemistry. The mathematical description of this
isotherm almost invariably involves either the Langmuir equation or the
'0n Bemmelen-Freundlich equation." The Langmuir efl,:"atio~.has the form
q

bKc
1 + Kc

(4.7)

where band K are adjustable parameters. The parameter b represents the


value of q that is approached asymptotically as c becomes arbitrarily large.
The parameter K determines the magnitude of the initial slope of the
isotherm. The most precise way to determine these two parameters with
experimental data is to plot the ratio
(4.8)
known as the distribution coefficient, against the surface excess, q.' After
multiplying both sides of Eq. 4.7 by 1/c + K and solving for K d , one finds
that the Langmuir equation is equivalent to the linear expression
Kd

bK - Kq

(4.9)

~
to
s~ould be

+ "'''
Thus a graph of K d ~ainst q
a straight line with slope - K and an
x-intercept equal to b if the LanewJ,!ir ~QyatiQn is appli"able. An example
of this kind of graph was presented in Fig. 1.12.
Not uncommonly, it is observed that a graph of K d against q is a curve
convex to the q-axis instead of a straight line. An example of this kind of
graph is shown in Fig. 4.2 for phosphate adsorption by a clay loam soil. 8 If
the value of K d tends to a finite constant as q tends to zero and if K d
extrapolates to zero at some finite value of q , then the adsorption isotherm
can always be fit to a two-term series of Langmuir equations:"
II

q-

blKlc
b
+ 2K2c
1 + Klc 1 + K 2c

(4.10)

where b l , b2 , Kl> and K 2 are adjustable parameters. This fact can be


illustrated by setting c = q I K d in Eq. 4.10 and clearing the fractions on
the right side to obtain the second-degree equation
K~

+ (K, + K 2)Kd q + K lK2q 2 - (blK l + b 2K2)Kd

bKlK2 q

=0
(4.11)

where b = b l + b 2 The derivative of K d with respect to q follows from


Eq. 4.11:

ss,

--=

dq

(K. + K 2)Kd + 2K.K 2 q - bK tK 2


2K.. + (K. + K 2)q - (b.K. + b 2K 2 )

(4.12)

119

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS


1.0
(

P SORPTION
WATTS SOIL
t = 25 IC

0.8

II>

C'

.><

<,

c:

0.6

o
..:;,

o
II>

0.4

'"E
"0

INTERCEPT = a~1 I a II

~ 0.2

SLOPE =fJO 1"'1

INTERCEPT=l3o

10

q (mmol PI kg soil)

Figure 4.2. A graph of the distribution coefficient against the amount adsorbed for
o-phosphate adsorption by a clay loam soil. 8 The parameters labeling the lines
through the data points are defined in Eqs, 4.13 and 4.14.

As q tends to zero, Eqs. 4.11 and 4.12 show that


K d = ao

+ (adao)q

(q ~ 0)

"(4. 13a)

where
ao = blK I + b 2K2

al =

(4.13b)
(4. 13c)

-(bIKi + b 2KD

Thus the distribution coefficient is linear in q near the origin, with a slope
equal to al/aO' According to Eq. 4.13a, the x intercept of the linear
expression is a5llall, as indicated in Fig. 4.2. On the other hand, when K d
extrapolates to zero, q = b according to Eq. 4.11. Equation 4.12 can then
be used to demonstrate that
(q

b)

(4. 14a)

where

f30

= b = bl

+ b2

-b l
b2
{31 = - K1
K2

(4.14b)
(4. 14c)

Thus the distribution coefficient again becomes linear in q as q tends to b,


its maximum value according to Eq. 4.10. The slope of the line is {Jo/{3t,
and its x intercept is, of course, b. If adsorption data are plotted as in

THE SURFACE CHEMISTRY OF SOILS

120

Fig. 4.2, then the limiting slopes and the two x intercepts can be determined graphically. The values found can be substituted into Eqs. 4.13b,
4.13c, 4.14b and 4.14c to solve uniquely for the Langmuir parameters b 1 ,
K 1 , b z, and K Z 9
The van Bemmelen-Freundlich isothe!!!! equation ha.s the form
q = Ac f3

(4.15)

where A and f3 are positive, adjustable parameters, with f3 constrained to


lie between 0 and 1. This expression can be derived by generalizing
Eq. 4.10 to an integral over a continuum of Langmuir equations.!"
q =

OO

()

m y

-00

exp(y)c
dy
1 + exp(y)c

(4.16)

where y = In K and m(y) is the weighting factor for the Langmuir term in
the integrand whose K parameter equals exp(y). This weighting factor is
subject to the constraint
b =

f~oo m(y)dy

(4.17)

where b is the maximum possible value of q, as before. As it is written,


Eq. 4.16 is completely general and can be used to derive any isotherm
equation by insertion of an appropriate choice for the weighting factor .10,11
For example, if m(y) is set equal to a sum of two delta "functions,"

m(y) = b 1 5(y - In Kd + b z5(y - In K z)


where b 1 , K 1 , b z , and K z are constant parameters, then Eqs. 4.16 and
4.17 become the same as Eqs. 4.10 and 4.14b, respectively. (The defining
relationship for the delta function,

f~oo f(x)5(x -

a)dx == f(a)

where a is a constant parameter, is used in the calculation.V) In order to


derive the van Bemmelen-Freundlich equation, m(y) must take the form:

m(y)

2 Cos(7Tf3)exp[f3(Ym - y)] + 2 exp[f3(Ym - y)]


mmax 1 + 2 Cos(7Tf3)exp[f3(Ym - y)] + exp[2f3(Ym - y)]
(4.18)

where
b
mmax = 27T tan( 7Tf3/2) = m(Ym)

Ym =

In (A/b)
f3

(4.19)
(4.20)

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOilS

121

Then it can be shown thatlO,ll


q

Ac 13
1 + (A/b)c13

(0 < (3 < 1)

(4.21)

In the limit of (A/b)cl3~ 1, Eq. 4.21 reduces to Eq. 4.15.


The weighting factor m (y ) is very similar to a gaussian function centered
on Ym, as illustrated in Fig. 4.3. Thus the van Bemmelen-Freundlich
isotherm can be thought of as the result of a log-normal distribution of
Langmuir parameters K (i.e., a normal distribution of In K) in a soil. The
parameter (3 determines the narrowness of this distribution, in thatlO,ll
liml3t1 m(y) = b 8(y - Ym)

(4.22)

follows from letting {3 approach unity in Eq. 4.18. As shown in Fig. 4.3,
the closer {3 is to zero, the broader is the distribution m (y ); the closer (3 is
to unity, the more m(y) approaches the delta function centered on Ym'
given in Eq. 4.22. In this limit, the introduction of the right side of
Eq. 4.22 into Eq. 4.16 reduces Eq. 4.16 to Eq. 4.7, the Langmuir equation, since the distribution is now just an extremely sharp-peaked spike
at Ym = In(A/b) == In K.
The weighting factor, m (y) in Eq. 4.18 can be calculated explicitly for
any soil if A, b, and (3 have been determined experimentally from
adsorption data. This estimation of parameters can be done by plotting log
Figure 4.3. A graph of the distribution of In K values, m(y) (y = In K), that leads
to the van Bemmelen-Freundlich isotherm: Each curve corresponds to the same
value of Ym but to different values of ~ in Eq. 4.18.
m(y)/b

{3 =0.9

"
,

x
t

................ ".
-'I

......................._....:.;-;.;1 ....

{3 = 0.8

{3-05

....,.,.:.:::.:-...:.......

---_

--

'1 m

.
+'1

122

THE SURFACE CHEMISTRY OF SOILS

q against log c for the range of concentrations over which Eq. 4.15 applies,
in order
c ulate log A and {3 from the y intercept and slope of the
resultin straigh ine. Then the variable q / c(3 is plotted against q to
defermine the value f A / b according to the expression
q/c(3 = A - (A/b)q

which is derived from Eq. 4.21 after both sides are multiplied by
c- (3 + (A / b) and the ratio q / c(3 is solved for. The x intercept for the linear
plot equals the parameter b. These operations emphasize the point that the
van Bemmelen-Freundlich equation applies strictly to adsorption data
obtained for low values of c.
4.2. ADSORPTION VERSUS PRECIPITATION

Fundamental to the interpretation of a loss of some substance from a soil


solution as an adsorption process is the hypothesis that the phenomenon
involved actually occurs on a surface. Adsorption is defined in Sec. 4.1 as a
net accumulation at an interface; precipitation can be defined as an
accumulation of a substance to form a new bulk solid phase. Both of these
concepts imply a loss of material from an aqueous solution phase, but one
of them is inherently two-dimensional and the other is inherently threedimensional. 13 However, the distinction between the two begins to blur
after one realizes that the chemical bonds formed in both can be very
similar, and that mixed precipitates can be inhomogeneous solids with one
component restricted to a thin outer layer because of poor diffusion. In ..
soils, the problem of differentiating adsorption from precipitation is made
especially severe by the facts that new solid phases can precipitate
homogeneously onto the surfaces of existing solid phases and that weathering solids may provide host surfaces for the more stable phases into which
they transform chemically. These conditions serve as a caveat against the
hasty application of Eq. 4.6 when only the loss of a substance from an
aqueous solution phase has been measured. When no independent data on
which to base a decision are available, this loss of material to the solid
phases in a soil can be termed simply sorption in order to avoid the
implication that either adsorption or precipitation is occurring.
The natural question to pose at this juncture is, what experimental
evidence is sufficient to determine whether a substance lost from the soil
solution has, in fact, been adsorbed? An answer to this difficult and
complex question can perhaps best be introduced by a discussion of
experimental criteria that cannot be used to distinguish adsorption from
precipitation in soils. This kind of discussion should be of help in any
critical reading of the abundant literature on sorption studies, as well as in
the design of experiments on surface phenomena in soils.
The adherence of experimental sorption data to an adsorption Isotherm
equation provides no evidence as to the actuaJ mechanism of a sorption process
In a soli.

'I
\,
:1

,
,j,
I

,I

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOilS

123

~ent can be validated both through model experiments and with

rigorous mathematical analyses of the properties of adsorption isotherm


equations.l" However, the point that these equations are not unique
mathematical descriptions of surface reactions in soils can be made
effectively by a simple counterexample. Consider a typical batch sorption
experiment in which o-phosphate is reacted with soil material and suppose
that in this reaction an amorphous aluminum phosphate phase is formed.P
It will be assumed that the conditional solubility product
)

applies to the amorphous solid phase, where AIT and P0 4T are the total
molar concentrations of aluminum and o-phosphate in the soil solution. If
n AI is the total number of moles of aluminum in the soil available to react
with added o-phosphate, then it is reasonable to expect that the following
inequality holds:
(4.23)
where V w is the volume of soil solution in cubic decimeters. Further, if the
concentration of added o-phosphate is large enough to convert most of the
available "reactive" aluminum in the soil to aluminum phosphate, then
(4.24)
where n AlPO. is the number of moles of phosphate precipitated. With the
definitions

where M; is the mass of solid soil material reacted with phosphate,


Eq. 4.24 can be written in the form
T

qpo. = b po . ( 1 - AI V
nAI

The use of
definition

cKso

w)

(4.25a)

to eliminate AIT from Eq. 4.25a and the additional

(4.26)
lead to the expression

(4.25b)
Finally, since

THE SURFACE CHEMISTRY OF SOILS

124

acco
the expression for cKso and the inequality 4.23, Eq. 4.25b can
De approxim ted closely with the inverse binomial expansion of the factor
in parentheses n the right side:
\.
b po 4P0
b po KP0 4T
QP04 = 1 + (1/ K
) = 1 / KP0 4T
4T

(4.25c)

Under the conditions described in Eqs. 4.23 and 4.24, the quantity of
a-phosphate precipitated is described by a Langmuir equation (Eq. 4.7)
even though surface reactions are not involved. Note that a graph of
nAJPO/P04T against nAJP04, based on the sorption data collected, leads to
a determination of nAl as the x intercept, in keeping with Eq. 4.9, and that
the "affinity parameter" K, calculated from the slope of a line cutting the
x axis, has no possible interpretation as a surface complexation equilibrium
constant. Instead, K is determined by the amount of reactive aluminum in
the soil, the soil water content, and cKso, as shown by Eq. 4.26. These
results have a direct bearing on batch sorption experiments designed with
relatively high concentrations of added a-phosphate as a means of estimating the maximum surface capacity of a soil for phosphate. If the added
phosphate is precipitated instead of adsorbed, Eq. 4.25 may apply and the
estimated capacity parameter has no particular relation with the phosphate-adsorbing surfaces in the soil.
The experimental observation that an ion activity product in a soil solution is
smaller than a corresponding solubility product constant provides no evidence
as to the actual mechanism of a sorption process in the soil.

This statement refers to a comparison between the ion activity product


(lAP),

lAP

= (Mm+)a(LI-)b

(4.27)

and the solubility product constant,


K so

(Mm+)a (Ll-t
(MaLb(s))

(4.28)

that pertain to the dissolution reaction


MaLb(s) = aMm+(aq) + b Ll-(aq)

(4.29)

In Eq. 4.28, M is a metal and L is a ligand that precipitate to form the solid
MaLb(s), ( ) is a thermodynamic activity, and a and b are stoichiometric
coefficients subject to the constraint of electroneutrality.P
am = bl

(4.30)

It is evident from Eqs. 4.27 and 4.28 that

lAP

."

= (MaLh(s

(4.31)

and therefore that the relation between lAP and K." (at fixed temperature

1NC)R(:1At"rrc-7\ND ORGANIC SOLUTE ADSORPTION IN SOILS

125

and pressure) is termined by the activity of the precipitated solid phase,


MaLb(s). An expe 'mental finding that lAP is less than Kso-often
interpreted as evidence for an adsorption reaction-may mean only that
the activity of the solid phase in the precipitate formed is less than unity.
The principal mechanism for a reduction in the activity of a solid phase
below its standard-state value of unity in soils is copreeipitation, which is
interpreted in chemical composition data as isomorphic substitution of one
ion for another of comparable radius in a soil mineral. It is illustrated in
Table 1.4 that isomorphic substitutions of cations cover a broad range of
both metals and minerals in soils. The same is true for inorganic anions,
and this fact is especially pertinent to investigations of the sorption
behavior of trace constituents of soil solutions, whose content in the solid
phases is expected to be small enough to preclude direct detection of their
presence in a precipitate. For example, if M is a trace metal that has
coprecipitated with a macroconstituent metal in a soil solution, then the
activity of the component containing M is likely to be quite low, decreasing
with the mole fraction of M in the mixture. 16 This reduction of the activity
of the solid component is accompanied by a corresponding reduction in the
activity of Mm+(aq), as can be seen by solving Eq. 4.28 for (M m+):

( M m+) '= [Kso(MaLb(S))]l/a


(L1 )b
= (M m+)O(MaLb(s))l /a

(4.32)

where
(4.33)
is the activity the metal cation would have in the soil solution if MaLb(s)
were in the standard state with unit activity.
Equation 4.32 shows that (M'?") is reduced below (Mm+)o when
(MaLb(s)) is less than unity. As a numerical illustration of this idea,
consider a calcareous soil to which a wastewater containing cadmium has
been applied, with the result that a coprecipitate of CdC0 3(s) and
CaC0 3(s) forms. Measurements on the aqueous phase of the soil indicate
that (Cd2+) = 10- 6 . 5 and (HCO)) = 10- 3 at pH 7.6. Because K so =
10- 1 1.2 for the reaction
CdC03(s) = Cd2+(aq)

CO~-(aq)

and (HC()t)7(C()~\) = 102 .67 at pH 7.6,17 the measured bicarbonate


activity and ECq-.~;31 produce the result
/'~

(Cd2+)O = 10- 1 1.2

,i

._____-n"
...

105 .67 ~(1O-5.53

It follows that the observed activity of Cd 2+(aq) can be interpreted with


Eq. 4.32 as an effect of the precipitation of a mixed Cd/Ca carbonate. with

THE SURFACE CHEMISTRY OF SOILS

126

the CdC0 3 component having an activity of about 0.1:


(Cd2+)

(CdC03(s))

= (Cd2+)o

_
= 10

The single fact that lAP = (Cd2+)(CO~-) = 10-12 . 2 is much smaller than
K so = 10-11.2 in this example cannot be used unambiguously to infer that
cadmium has been adsorbed by the soil. Both ~4so,rptiona9d precipitation
are consistent with an lAP diminished below the solubility product
constant for a possible solid phase.
The experimental observation that an ion activity product in a soil solution is
larger than a corresponding solubility product constant is not prima facie
evidence of precipitation.

If a solid phase has precipitated according to the reverse of the reaction

in Eq. 4.29, the lAP can be larger than the corresponding solubility product constant for the solid if the activity of the solid is greater than unity
(Eq. 4.31). This circumstance occurs commonly when the precipitate
comprises particles whose radii are smaller than about 1 JLm. 16 The surface
energy of these very small particles is large enough to contribute importantly to the Gibbs energy of the precipitate and therefore to increase its
activity relative to that in the standard state, where the interfacial energy
component of the Gibbs energy is negligible by definition. On the other
hand, the simple condition of supersaturation, lAP/ K so > 1, is not sufficient. for the actual formation of a solid phase at a measurable rate. The
rate of homogeneous precipitation depends, for example, sensitively on
the degree of supersaturation such that, under circumstances typical of soil
solutions, solid formation requires geological time intervals if lAP/ K so is
smaller than about 20. 16 If nucleating agents are present in a soil, the rate
of heterogeneous precipitation from a supersaturated soil solution should
be large and the mechanism of sorption should be solid formation, not
adsorption. This mechanism is especially probable when solid phases that
are similar to the one expected are already present in a soil, as when
carbonate-forming trace metals are introduced at supersaturation levels
into the aqueous phase of a calcareous soil.
It should be clear at this point that classical thermodynamic methods
involving adsorption isotherm equations and the solubility product principle cannot distinguish adsorption from precipitation unambiguously. This
fact is just another illustration of the impossibility of inferring underlying
mechanisms from thermodynamic data on soils. Of course, a relatively
complete determination of the chemical composition of a soil solution after
sorption has occurred can provide useful circumstantial evidence. For
example, a measurement of the concentration of silicon in a soil solution is
essential if the sorption of an oxyanion by a soil is under investigation,
since the precipitation of an oxyanion through the incongruent dissolution
of aluminosilicates is accompanied by the release of silicon. As a general

~..
INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

127

rule, the chemical analysis done after a sorption experiment with soil
should include all possible relevant elements in the soil solution, not just
the one that is the chief object of investigation.
J
Perhaps the best methods for demonstrating the existence of adsorption
in a soil are optical, magnetic resonance, and X-ray photoelectron spectroscopy, which give direct evidence for the presence of adsorbed species. These methods currently are under development for application
to soils extensive calibration with well-characterized, reference soil
minerals. is Until this calibration is completed, it is possible to use kinetics
data to make an operational distinction between adsorption and precipitation. This strictly empirical method of analyzing sorption data can be
illustrated with the important case of o-phosphate reactions.
It has been recognized for about 40 years that the reaction between
o-phosphate and soil exhibits rapid and slow stages.l" The rapid stage
almost invariably persists for less than 50 hours, and the slow stage
continues well beyond 50 days in many instances. No particular mechanism
of phosphate sorption can be inferred uniquely from the kinetics data that
show these two stages, but it is not unreasonable to suppose that, if the
initial concentration of phosphate in the soil solution is below supersaturation for any possible phosphate solid, the rapid stage corresponds
principally to adsorption. Besides the expectation that a surface reaction
should proceed quickly in the absence of diffusional barriers, supporting
evidence for this hypothesis comes from the fact that rapidly sorbed
phosphate is also readily desorbable.l?
On the other hand, a classification of a sorption process on the basis of
kinetics data must be conditioned by other chemical properties of thFphosphate-soil mixture. For example, if the soil solution is supersaturated
initially with respect to some phosphate solid, precipitation is likely to
influence the sorption reaction from the beginning.F" If the soil minerals
have a low degree of crystallinity and/or a high degree of hydration,
precipitation may be the dominant sorption mechanism even in the rapid
stage." In general, low phosphate concentrations and well-crystallized,
relatively unhydrated soil minerals tend to favor adsorption as the phosphate reaction mechanism. Other chemical properties, such as the pH
value of the soil solution and the kinds of metals in soil clay minerals, exert
a quantitative influence on the rapid stage of phosphate sorption, as do
such physical properties as temperature. 22
A large number of mathematical expressions have been applied to
describe the rate of phosphate sorption in soils, but no clear consensus on
which equations are most suitable has yet emerged. 22 ,23 There is at present
a growing use of the Elovich equation'" to represent the rapid stage:
dq

dt = k, exp( -k2q )

(4.34)

where q is the amount of phosphate sorbed per unit mass of dry soil and k,
and k2 are constant parameters. Although Eq. 4.34 can be derived from

THE SURFACE CHEMISTRY OF SOILS

128

more general rate-law expressions under the assumption of an exponential


decrease in number of available sorption sites with q and/or a linear
increase in activation energy of sorption with q,25 the Elovich equation is
perhaps best regarded as an empirical one for the characterization of rate
data. The solution of Eq. 4.34 is: 25

t)

q(t) = k1 In(k 1k2to) + k1 In ( 1 + to

(4.35)

where

to =

exp(k 2 qc)
k

1k2

tc

(tc

>

0)

and qc is the value of q at time tc, the time at which the rate of sorption
begins to be described by Eq. 4.34. In most applications, tc is set equal to
zero and qc is the initial value of q in the soil. Equation 4.35 appears to
describe phosphate sorption by soils quite well for t < 200 hours.r" Some
evidence exists to suggest that, over this period, phosphate sorption
involves principally a multilayer adsorption mechanism, i.e., the formation
of metal phosphate coatings on the surfaces of soil particles.P This
hypothesis has the attractive feature that it is consistent with the ultimate
formation of a precipitate in the slow stage of phosphate sorption. The
latter appears to be described well by the linear rate law26
dP0 4T
dt
= - K P0 4T

4.36

where P0 4T is the total molar concentration of phosphate in the aqueous


solution phase and K is a constant parameter. Equation 4.36 has been
shown to apply for times of reaction longer than about 140 hours and to be
associated with the appearance of discrete crystallites of phosphate
solids.r" Evidently the transition from monolayer adsorption to multilayer
adsorption to precipitation is marked by a change in the rate law from
Eq. 4.34 to Eq. 4.36. As a rule of thumb, the adsorption-dominated stage
of phosphate sorption can be assumed to exist for reaction times less than
about 50 hours.
4.3. METAL CAliON ADSORPTION

The positive adsorption of metal cations by the solid phases in soil can
involve the formation of either inner-sphere or outer-sphere surface
complexes, or the simple accumulation of an ion swarm near the solid
surface without complex formation. These adsorption mechanisms are
implied in the development of the concept of surface charge balance
(Eq. 3.3) and were illustrated, for the case of surface complex formation,
in Figs. 1.8 and 1.10. The quantitative relationship between these mechanisms and measured surface excesses of metals on soil minerals is taken
up in Chap. S. In the present section, emphasis is placed on the qualitative

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

129

aspects of metal cation adsorption in soils ois-a-ois surface complexation


phenomena.
An important leitmotiv throughout the long
!!!story of metal cilJ:i2!!..~sOrI?ti<2!!E.!:Is!i~.L~l~1liU'.tt-t~.!:!1q.~.
27
This preoccupation with the relative affinity of soil particle surfaces for
metal cations is reflected in the abundant literature on cation exchange
selectivity coefficients and the frequent published attempts to deduce
"replaceability series" from experimental adsorption data. 28 It is an
unfortunate fact that, in most studies, no effort has been made either to
define affinity and selectivity precisely or to exclude from the experiments
the confounding effects of multiple cation adsorption and soluble complex
formation between the metal cations under investigation and the ligands in
the soil solution. The result of these common oversights has been the
emergence of a variety of selectivity sequences, with no particular uniformity even among those deduced for a single, well-characterized soil
mineral. 28,29 These difficulties are least troublesome for monovalent metal
cations, however, because the different definitions of selectivity based
either on chemical thermodynamics or on a direct comparison of adsorption isotherms coincide when only monovalent cations are involved and
because the extent to which these cations form soluble complexes in a soil
solution is usually very small.i" Accordingly selectivity sequences for the
adsorption of monovalent cations by soil particles are the most easily
interpreted mechanistically.
The smectite minerals offer perhaps the clearest opportunity to examine
how surface complexation by the siloxane ditrigonal cavity affects monovalent cation adsorption. Since it is quite reasonable to assume that there will
be no differences in monovalent cation selectivity on smectite surfaces if
the only mechanism of adsorption is the accumulation of a cation swarm in
the soil solution near the solid-liquid interface." the existence of a
selectivity sequence for smectites must imply the formation of surface
complexes, such as those illustrated in Fig. 1.8. If the observed differences
in monovalent cation selectivity are principally the result of inner-sphere
complex formation (i.e., a monovalent cation settled into a siloxane
ditrigonal cavity), then the expected selectivity sequence should be
Cs+ > Rb+ > K+ > Na+ > Li+ for the Group IA metal cations.P The
molecular basis for this sequence can be understood as follows. Innersphere complex formation with a siloxane ditrigonal cavity requires the
exchange of some of the water molecules solvating the adsorptive cation
for the oxygen ions that constitute the surface functional group. As
discussed in Sec. 1.2, these oxygen ions are relatively soft Lewis bases
when compared with the oxygen atoms in water. It follows from the
principle of hard and soft acids and bases (HSAB) that the most stable
inner-sphere complexes involving the siloxane ditrigonal cavity form with
the Group fA cations that are the softer Lewis acids because these cations
more readily exchange their solvation water molecules for a softer Lewis
ADSORPTION SELECTIVITY.

lin SUIUAU CHfMISTRY Of SOILS

130

base.:" The relative Lewis acid softness of a Group IA cation can be


estimated numerically with the Misono softness parameter, y:34
lOIzr

y = ---=::...;=
I Z +1

(4.37)

JZ

where I z is the ionization energy for a cation of valence Z and radius r. For
the Group IA cations, Z = 1 and the sequence of decreasing Y values is
Cs+(O.287) > Rb+(O.228) > K+(O.189) > Na+(O.lll) > Li+(O.053)
(4.38)
where Y is in units of nanometers. This sequence is identical with that of
decreasing stability of inner-sphere complexes formed with the siloxane
ditrigonal cavity through the elimination of solvation water molecules.
Figure 4.4 shows values of the standard Gibbs energy, ~G~x, at 298 K
for the reaction
LiX(s) + M+(aq)

= MX(s) + Li Taq)

(4.39)

Figure 4.4. The relation between the standard Gibbs energy for Li + - M+
exchange on montmorillonite and the Misono softness parameter ofM+ (M = Na,
K, Rb, CS).35
12
Cs+

Li+ - M+ EXCHANGE
ON MONTMORILLONITE

10

Camp Berteau
Wyoming

Rb+

10
0

'I;

'i,i,"

'I

~'I

':,/j

-'"
...J

OQl

.~

K+

I
' ,~

II

<.!l

'I

<J

',I

'I
I

',I

Na+

',I

0.05

0.10
YM+

0.15

- YL1+(nm)

0.20

0.25

INOKGANIC AND ()K(;ANIC SOllJ II ADSOKI'1I0N IN SOli S

III

plotted against the difference in Misono parameters, Y M' - Y Li " for


M+ = Na+, K+, Rb+, and Cs+ on montrnorillonite.P (In Eq. 4.39, X
refers to one mole of negative montmorillonite charge.) The standard
Gibbs energy for the exchange reaction in Eq. 4.39 is a quantitative
measure of cation selectivity. It is evident in Fig. 4.4 that aG~x decreases
as the Misono parameter of M+ increases relative to the value for Li+. This
trend is in agreement with what is observed typically for smectites32.33.36
and with the sequence in Eq. 4.38. For a given value of (YM+ - YLi+).
Fig. 4.4 also shows that aG~x is smaller (i.e., the clay is more selective for
M+) for Camp Berteau montmorillonite than for Wyoming montmorillonite. These two smectites differ in that all of the negative charge on the
siloxane surface of the Camp Berteau montmorillonite originates from
isomorphic cation substitutions in the octahedral sheet, whereas only about
two thirds of the negative charge on the siloxane surface of Wyoming
montmorillonite originates in the octahedral sheet, with the remainder
produced by isomorphic cation substitutions in the tetrahedral sheet.
According to the discussion in Sec. 1.2, this difference in isomorphic
substitution makes the siloxane ditrigonal cavity on Camp Berteau montmorillonite a softer Lewis base than that on Wyoming montmorillonite. It
follows from this conclusion and the HSAB principle that, as the Lewis
acid softness of the complexed cation increases, Camp Berteau montmorillonite should form more stable inner-sphere complexes than Wyoming montmorillonite.
The hypothesis of inner-sphere complexation as the basis for differences
in monovalent cation selectivity on siloxane surfaces is consistent with the
observed decrease in SE, the specific surface area measured by chloride
exclusion, with an increase in Lewis acid softness of the exchangeable
cation, shown in Table 1.8. It is pointed out in Sec. 1.4 that surface
complexation between siloxane ditrigonal cavities and monovalent cations
reduces the negative charge that repels chloride ions from a siloxane
surface. Additional, indirect evidence supporting this view comes from
the results of Na" -+ Cs" exchange experiments with reduced-charge
montmorillonites.P As the negative charge on the siloxane surface of
Camp Berteau montmorillonite is reduced permanently through lithium
migration into the octahedral sheet at high temperature (the HoffmannKlemen method), the value of aG~x for the exchange of Na+ by Cs"
increases monotonically toward zero. This trend can be interpreted as the
effect of an overall increase in the Lewis base softness of the siloxane
surface induced by the decrease in surface charge density and a concomitant further delocalization of the negative charge produced through
isomorphic substitution. Since both Na + and Cs" actually are hard Lewis
acids (with Cs" the softer of the twO),33 an increase in the Lewis base
softness of the siloxane surface tends to decrease inner-sphere coinplex formation for both cations, according to the HSAB principle.
This decrease should be accompanied by enhanced dissociation of
both cations from the siloxane surface and therefore an increase in
aG~x for Na + - Cs+ exchange toward the zero value. indicative of no

132

THE SURFACE CHEMISTRY OF SOILS

thermodynamic preference for either cation by the clay mineral, as occurs


when the only adsorption mechanism is accumulation of an ion swarm in
the aqueous solution phase near the adsorbing surface.
Comparable statements concerning the basis for selectivity differences
among bivalent metal cations adsorbed by siloxane surfaces cannot be
made with certainty for two primary reasons. First, because of quasicrystal
formation, the extent to which outer-sphere complexes form between
bivalent metal cations and siloxane ditrigonal cavities is quite pronounced
and the degree of dissociation of these cations from a siloxane surface is
correspondingly very small. It follows that selectivity differences among
metals in oxidation state II are likely to be small, even if inner-sphere
complex formation is the causative factor. This expectation appears to be
in agreement with bivalent cation exchange data on smectites and
vermiculites.Pr" A second complicating difficulty is the fact that there are
almost no reports of experiments on bivalent cation exchange in which
soluble, complex-forming ligands were excluded from the aqueous solution
phase. Without this precaution, the measured surface excess and selectivity
coefficient must reflect the adsorption of both free bivalent cations and
their complexes and no unambiguous inference concerning the role of
surface complexes can be made.?? Despite these problems, it can be stated
that inner-sphere surface complexes do form between bivalent metal
cations and siloxane ditrigonal cavities, as discussed in Sec. 2.3, and that
the possibility of their having a role in determining differences in adsorption affinity remains open.
Similar difficulties beset the interpretation of data on metal cation
adsorption by surfaces bearing inorganic hydroxyl groups. Besides the
paucity of data on adsorption from aqueous solutions free of metalcomplexing ligands, the fact that the hydroxide ion is a much harder Lewis
base than the oxygen ion in a siloxane ditrigonal cavity must be considered.
Since both OH- and H 2 0 are very hard Lewis bases, a pronounced
tendency for the hard Lewis acids in Groups IA and IIA of the Periodic
Table to desolvate and form complexes with surface hydroxyl groups as the
Lewis acid softness of the metals increases are not likely. In this case, more
subtle effects relating to the distribution of surface charge (e.g., the
Brensted acidity of the surface hydroxyl group) may determine the extent
of inner-sphere complex formation. The selectivity sequences reported for
monovalent and bivalent cation adsorption by hydrous oxides support this
hypothesis, in that strongly (Brensted) acidic surfaces tend to show
increasing selectivity with increasing Lewis acid softness, whereas weakly
acidic surfaces show the opposite trend. 29
The effects of metal-complexing ligands in a soil solution
on the adsorption of metal cations by soil constituents can be classified into
four general categories;"
LIGAND EFFECTS.

1. The ligand has a high affinity for the metal and forms a soluble complex
with it, and this complex has a high affinity for the adsorbent.

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

133

Figure 4.5. A schematic diagram of interactions in an aqueous metal-ligandsurface system. The parameters K refer to equilibrium constants for reactions in the
direction of the arrows.

2. The ligand has a high affinity for the absorbent and is adsorbed, and the
adsorbed ligand has a high affinity for the metal.
3. The ligand has a high affinity for the metal and forms a soluble complex
with it, and this complex has a low affinity for the absorbent.
4. The ligand has a high affinity for the adsorbent and is adsorbed, and the
adsorbed ligand has a low affinity for the metal.
A fifth category could be the ligand that has a low affinity for both the
metal and the adsorbent and therefore little or no effect on trace metal
adsorntlOn (e.g., clu 4 at pH > PZC for the adsorbing solid). The four
main categories can be deduced from the overall scheme of metal-ligandsurface interactions depicted in Fig. 4.5, as shown in Table 4.1. The
scheme of interactions emphasizes the competition between the metal and
the adsorbent for the ligand. Categories 1 and 2 in Table 4.1 should result
in enhanced adsorption of the metal. If the ligand and metal do not interact
with the same surface functional groups, category 4 produces little effect on
metal adsorption; if there is competition, metal adsorption is reduced.
The model systems listed in Table 4.1 represent well-characterized
metal-ligand-adsorbent combinations whose observed behavior is consistent with at least one category of ligand effecta" On kaolinite, a sharp
Table 4.1. Categories of ligand effects on metal cation adsorption
Category
1
2

Equilibrium constant relations

K ML K ML X large; K OM
K L X K M/ LX large; K OM
K ML large; KMl. X small

K L X K Il M

large;

KM/LX

K OL small
KOL small

small

Model system example.'?


Cu-hydroxide-kaolinite
Cu-glutamate-Fe (OHM am)
Cu-glycine-rnontmorillonite
Cu-phosphate-illite

134

THE SURFACE CHEMISTRY OF SOILS

increase in the adsorption of copper is produced by a change in pH from


5.5 to 6.0, suggesting the adsorption of Cu-hydroxy complexes by the clay
mineral. On amorphous iron oxyhydroxide, glutamic acid is expected to be
complexed through one of its carboxyl groups to a surface hydroxyl group,
leaving a zwitterion exposed to the aqueous solution phase. The carboxylate in the zwitterion can then complex a Cu2+ ion. On montmorillonite,
the adsorption of copper is reduced through the formation of monovalent
and neutral Cu-glycine complexes that do not react well with the negatively
charged clay surface. On illite, the adsorption of phosphate is thought to
involve principally aluminol groups, which are not significant in copper
adsorption by the clay mineral. The adsorbed phosphate evidently is not
effective in forming complexes with Cu2+ ions in the aqueous solution
phase, since no enhancement of copper adsorption is observed. However,
the situation can be different when the adsorbent is a hydrous oxide instead
of a phyllosilicate. In this case, oxyanions such as o-phosphate can react
strongly with the surface hydroxyl groups to accumulate in multilayers that
in turn act as a new adsorbent for metal cations.t"
Perhaps the most important soil solution ligand that affects metal cation
adsorption is the hydroxide ion. Figure 4.6 shows four representative
examples of the effect of increasing OH- activity on metal adsorption by
inorganic and organic surfaces of soil constituents. 41 Data pertaining to this
effect are almost always summarized in a graph that pairs some quantitative measure of the extent of adsorption with the pH value of the aqueous
solution phase. In each of the graphs in Fig. 4.6, the adsorptions of metals
from solutions containing a given total metal concentration are compared
at different pH values. The curves show that increasing the OH- activity
always increases the extent of metal adsorption. Therefore, OH- is a
ligand whose effects on adsorption fall into categories 1 and 2. At present,
there are different opinions about which category is the more appropriate.
For example, it is possible that hydrolytic species of metals exhibit a very
high affinity for proton-selective surface functional groups, such as those
on oxide minerals. Alternatively, it can be proposed that OH- is adsorbed
by proton-selective surface functional groups with great affinity and then
serves as a bridge between the adsorbent surface and an adsorptive metal
cation.
It is often observed that metal cation adsorption data obtained over a
range of pH values can be described mathematically with the equation'?

= a + bpH

(4.40)

_/sorb
D -- K d C --

(4.41)

In D
where

Isoln

K d is the distribution coefficient, c, is the (fixed) mass of adsorbent per unit


volume of aqueous solution phase, I."rh is the ratio of moles of metal
adsorbed to (fixed) moles added initially, and 1."ln = 1 - I..,rh is the

135

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

a-FeOOH
!AT =10-4M
O.IM KN03

8
120r-----,--,..-----,--,---,
Cu

Pb

Co

...
<,

'0
E 80

::l..

~
u

c
0=-3---:-----:----:---::-------:

8
pH

pH

Figure 4.6. The effect of increasing OH- activity on the adsorption of metal
cations by solid soil constituents."

fraction of initially added metal that is not adsorbed. The parameter D is


called the distribution ratio and a graph of In D against pH according to
Eq. 4.40 is termed a Kurbatov plotY
A chemical interpretation ofthe constant parameters a and b in Eq. 4.40
can be made as follows. The pH value at which half of the added metal is
adsorbed is designated pH so . Since D = 1 at this pH value, according to
Eq. 4.41, it follows from Eq. 4.40 that
pH so

-a

Moreover, after substituting fsorb/(l - fsorb) for D in Eq, 4.40, one can
derive the result
dflOrb)
= b
( dpH pH-pH", 4

(4.42)

136

THE SURFACE CHEMISTRY OF SOILS

Therefore, Eq. 4.40 can be rewritten in the form


In D = b(pH - pHso)

(4.43a)

Isorb = {l + exp[ - b(pH - pHSO)]}-l

(4.43b)

or

with b as expressed in Eq. 4.42.


The values of pHso for a series of bivalent metal cations at the same
initial molar concentration, reacted with the same hydrous oxide adsorbent
suspended at the same solids concentration in the same background
electrolyte solution offer a relative measure of the selectivity of the
adsorbent for the metal cations. 29 The smaller the pHso , the more selective
the adsorbent for the metal cation. With a given metal, pHso is usually well
below the negative common logarithm of the stability constant for the
formation of the complex MOH+(aq) and often decreases as the initial
molar concentration of the metal decreases, unless the quantity of metal
adsorbed is far below the maximum possible, in which case pHso often
remains independent of the initial metal concentration.r" The slope
parameter b in Eq. 4.43 provides a measure of the steepness of the
"adsorption edge" at pHso (Fig. 4.6). This characteristic of a plot of Isorb
against pH cannot be attributed to anyone feature of the metal adsorption
mechanism, but the sign of b can be used to classify the plot as either
"metal-like" (b > 0) or "ligand-like" (b < 0) when metal-complexing
anions besides OH- are present. 38,44 (These terms are compared further in
Sec. 4.4.)
Further insight as to the effect of hydroxide ion on the mechanism of
metal cation adsorption by surfaces bearing hydroxyl groups can be
obtained from the results of electrophoretic mobility measurements. It has
been known for about 40 years that the electrophoretic mobility of an
oxide or phyllosilicate particle can be made to change sign as the pH value
is increased in the presence of a sufficiently high concentration of bivalent,
trivalent, or tetravalent metal cations.P An example of this behavior is
shown in Fig. 4.7 for particles of Ti0 2 (rutile) suspended in solutions of
Ca(N03 h .46 The minima in the plots of u against pH that appear as the
concentration of calcium is increased are typical. For metal cations that
form more stable hydroxy complexes than Ca2+ (e.g., C0 2+), u exhibits a
maximum in the high-pH region and declines thereafter back toward
negative values.f? Thus the protoypical mobility-pH curve at sufficiently
high adsorptive metal concentrations simply tracks the mobility-pH curve
determined at very low (or vanishing) metal concentrations at low pH
values, then shows a minimum and reversal of the sign of the mobility at
intermediate pH values, and finally exhibits a maximum followed by
decline as pH is increased further. The declining portion of the curve at
high pH values lies close to the electrophoretic mobility-pH curve of the
hydroxide solid formed by the metal cation. 47

137

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

Rutile in Co (N03)2
Fuerstenou et 01. (198()

Car
1.67 x 10- 3 M

>

l/)

C\J

-
E

00

1.67x1O:-5M

:J

-I

10

II

pH
Figure 4.7. The dependence of the electrophoretic mobility of rutile particles on
the pH value in solutions of Ca(N03h .46

These characteristics of the electrophoretic mobility-pH relationship


suggest that metal cation adsorption by oxide surfaces (including the edge
surfaces of phyllosilicates) is accompanied by the formation of a metal
hydroxy-polymer coating on the adsorbent. The extent of this coating
depends on the pH, on the initial concentration of metal cations, and on
the ionic potential of these cations, along with other factors that determine
their hydrolysis. Ionic potential (the ratio of valence to ionic radius) is an
especially important property, in that metal cations with ionic potentials
between 30 and 100 nm -1 (e.g., Al (III) ) can be expected to form extensive
coatings readily, even on siloxane surfaces (Fig. 1.11), whereas metal
cations with ionic potentials less then 30 nm ? (metals of Group II in the
Periodic Table) do so only at relatively high pH values. However, it does
not seem possible to ascribe the formation of hydroxide coatings to an
ordinary precipitation reaction involving the metal cation. The observed
values of pH~lI for metal adsorption on hydrous oxides are well below the

138

THE SURFACE CHEMISTRY OF SOilS

pH range of homogeneous precipitation, as is the pH value at which the


minimum occurs in electrophoretic mobility-pH curves. 29 ,4S In addition,
the frequently observed shift of pHso toward lower values as the initial
concentration of metal is decreased is just the opposite of what one would
expect if a precipitation reaction were taking place.r"
Metal cation adsorption by the solid phases in soils should be governed
by some combination of the surface complexation mechanisms deduced
from experiments on well-characterized minerals and organic solids.
In soils whose clay fraction is dominated by 2: 1 phyllosilicates, metal cation adsorption should involve principally both inner- and outer-sphere
complexes with siloxane ditrigonal cavities, although highly selective
adsorption of bivalent trace metals may occur through complexes with
edge-surface hydroxyl groups. In soils whose clay fraction is dominated by
1: 1 phyllosilicates and hydrous oxides, metal cation adsorption should
involve the formation of hydroxy-polymer coatings whose extent of surface
coverage depends sensitively on the Lewis acid strength (ionic potential) of
the metal cation as well as on the pH value of the soil solution. In soils
wherein organic surfaces are important to metal adsorption, it is likely that
both surface complexation by organic functional groups (see Table 1.5)
and hydroxy-polymer formation (Fig. 4.6) are involved. Superimposed
on these basic mechanisms are the effects of metal-complexing ligands
summarized in Fig. 4.5 and Table 4.1. These effects and the multiplicity of
inorganic and organic ligands expected typically in soil solutions combine
to make the detailed, quantitative description of metal cation adsorption in
soils a formidable task.

4.4. INORGANIC OXYANION ADSORPTION


The ionic potentials of the nonmetal elements in Groups IlIA to VIA
exceed 100 nm ? for their common oxidation states, and therefore these
elements form oxyanions instead of hydrolytic species in soil solutions. 49
The same tendency is observed for the Group VIB metals chromium and
molybdenum. Examples of inorganic oxyanions commonly found in the
aqueous phases of soil are B(OH)i, CO~-, N03", H 3SiOi, PO~-, SO~-,
AsO~-, SeO~-, MoO~-, and, when the oxyanion is multivalent, some of
the protonated forms. The qualitative and mechanistic features of the
adsorption of these oxyanions-a topic on which there is an abundant
literaturesO- s2-are the principal concerns of the present section. Quantitative models of inorganic oxyanion adsorption are described in Chap. 5.
A consensus exists that inorganic oxyanion adsorption by soil minerals
involves almost universally the two-step ligand exchange reaction
SOH(s)

+ H''{aq)

= SOH:t(s)

SOH:t(s) + L/-(aq) = SL1-/(S) + HzO(l)

(4.44a)
(4.44b)

where S is a metal cation, SOH(S) is one mole of inorganic surface

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

139

hydroxyl groups, and L l - is an inorganic oxyanion of valence l. The


protonation step (Eq. 4.44a) is thought to render the surface hydroxyl
group more exchangeable; if the concentration of the oxyanion in the soil
solution is sufficiently high, this step may not be necessary.P On the other
hand, the exchange reaction (Eq. 4.44b) does not occur readily until very
high concentrations of nitrate are reached in soil solutions. 51 The situation
with carbonate and sulfate anions may be intermediate in that these
oxyanions may sometimes adsorb by forming an outer-sphere surface
complex, like nitrate usually does, with the protonated hydroxyl groupr'"
(4.44c)
and sometimes engage in ligand exchange instead. 51 The reactions of
oxyanions with phyllosilicates appear to follow Eq. 4.44 as much as do
those with metal oxides in soils. In the case of the phyllosilicates, the
reactive surfaces are both the edge surfaces of broken crystallites and the
exposed surfaces of adsorbed hydroxy polymers. 51,55
Much experimental evidence has been adduced in support of the ligand
exchange mechanism for oxyanion adsorption. 50-52 The principal lines of
reasoning that lead to this mechanism can be summarized as follows.
1. The maximal oxyanion adsorption by a soil mineral under a prescribed
set of conditions depends on the pH of the soil solution. If only one
species of oxyanion exists, a graph of the maximal adsorption versus pH
value (the "adsorption envelope") exhibits a broad peak at a pH value
near the pKa for the conjugate Brensted acid, as illustrated in
Fig. 4.8. 56 If several protonated species of an oxyanion exist, the
adsorption envelope usually shows a monotonic decline with increasing
pH value, the decline becoming sharper as the valence of the principal
oxyanion species decreases (Fig. 4.8). These trends are consistent with
Eqs. 4.44 in that they suggest decreasing adsorption as the protonation
of the surface decreases but increasing adsorption as the valence of the
oxyanion decreases. With B(OH)4' the decrease in positive surface
charge and the increase in concentration of the single oxyanion species
with increasing pH value are not comparable above pH 7 and a peak in
the adsorption envelope is observed. For oxyanions with multiple
species, the small differences in adsorption affinity among species leads
to an adsorption envelope more conditioned by surface protonation
than by oxyanion valence. In agreement with this point of view is the
fact that arsenate and phosphate adsorption data on oxyhydroxides can
be renormalized and correlated successfully with the PZC of these
adsorbents." The characteristic adsorption envelopes in Fig. 4.8. are to
be contrasted with the "adsorption edge" curves in Fig, 4.6. It is
evident that the ligand-like behavior in Fig, 4.8, which is indicative of
decreasing anion valence and surface charge, is the complement of the
metal-like behavior in Fig, 4,6, which is indicative of decreasing cation
valence and surface charge.

140

THE SURFACE CHEMISTRY OF SOILS

2000~

10'
oX

COO
PhOSPhat~
a-FeOOH

0
00

150

E
E

0
00

0
0

"0
Q)

.a 100

(ArSenate
AI(OH)3(am)

L-

(/)

"0
0

c:
0

MA

A At:A

50

c:

Borate~

<t

Fe(OH)3(am) -

10

12

pH
Figure 4.8. Adsorption evelopes for phosphate on goethite, arsenate on amorphous aluminium hydroxide, and borate on amorphous iron hydroxide. 56 The .
ordinate values should be multiplied by ten for the arsenate data.

2. The rate of 3Zp exchange between dissolved o-phosphate and phosphate adsorbed on goethite or gibbsite follows the Elovich equation
(Eq. 4.34).58 The parameter k z in Eq. 4.34 is observed to be a function
only of temperature, whereas k 1 depends on the temperature and a
fractional power of the proton concentration in the aqueous solution
phase. The latter parameter also includes a first-order rate dependence
on the concentration of phosphate surface complexes. The dependence
of k 1 on a fractional power of the proton concentration reflects the
effect of the protonation reaction in Eq. 4.44a, and the first-order
dependence on the adsorbed phosphate concentration is expected if
the rate-determining step in 3Zp exchange is the breaking of either the
metal-phosphate bond in an inner-sphere surface complex or the
metal-oxygen bond in the adsorbent. 58,59 Thus the kinetics data for 3Zp
exchange, like the pH effect on the maximal adsorption of phosphate,
are consistent with the ligand exchange mechanism.
3. The rate of hydroxyl exchange on -Cr-O, is known to be orders of
magnitude lower than that on goethite (a-FeOOH).6o The rate and
extent of nitrate adsorption by these two reference minerals in suspension are essentially the same, with equilibrium attained in minutes. This
fact suggests that nitrate adsorbs according to outer-sphere complex
formation (Eq. 4.44c), not ligand exchange (Eq. 4.44b). Sulfate
adsorption by the two minerals is also similar in rate and extent;

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

141

however, the PZSE of the minerals is shifted upward in a sulfate


background solution. These results illustrate the "intermediate" surface
complexation behavior of sulfate, already mentioned above. The
adsorption data are consistent with outer-sphere complex formation,
but the PZSE shift upward could reflect inner-sphere complex formation, as discussed in Sec. 3.2. By contrast, the rate and extent of
phosphate adsorption by the two minerals are markedly different, the
rate being much lower and the extent of adsorption much larger on
goethite. This difference is consistent with a ligand exchange mechanism for phosphate adsorption by goethite and outer-sphere complex
formation on a- Cr2 0 3 ' 60
4. The infrared spectra of evacuated or moist metal oxyhydroxide films
. bearing adsorbed phosphate and other oxyanions have been interpreted
in support of a ligand exchange mechanisrn.W'' Perhaps the best
evidence comes from studies on goethite, in which the. decrease in
absorbance of the reactive A-type surface hydroxyl (Fig. 1.9) can be
monitored readily as a function of extent of oxyanion adsorption. Band
assignments in the infrared spectrum of phosphated goethite have been
used to propose the inner-sphere surface complex structure illustrated
in Fig. 1.9. 61 However, the spectroscopic approach is not conclusive in
this respect since not all of the bands expected for the surface complex
can be observed and those that are occur where absorptions are
found for a-phosphate in a variety of coordination environments
(Table 4.2).61,62 Added to this ambiguity is the possibility that oxyanion
complexation mechanisms can depend sensitively on the concentration
of the adsorptive anion in the aqueous solution phase, such that data
pertaining to evacuated films or even moist ones can be misleading in
relation to what happens in aqueous suspensions or wet soils. As the
example of nitrate adsorption at high concentrations shows.i" drying an
adsorbent in the presence of an oxyanion can promote ligand exchange
well out of proportion to its importance as an adsorption mechanism
under the conditions expected in natural soils.
Although no one of the above pieces of experimental evidence is
conclusive in itself, the group taken together constitute a powerful
argument for the applicability of Eq. 4.44b to oxyanion adsorption by soil
minerals. Other indirect evidence, such as the low rate and extent of
desorbability of phosphate versus nitrate,50 only tends to support this
conclusion. In a natural soil, the reactive inorganic surface hydroxyls exist
in a variety of adsorbents exhibiting a broad range of chemical composition
and crystallinity, This fact suggests that the activation energy for oxyanion
adsorption is likely to take on a continuum of values instead of being
monolithic. That the Elovich equation (Eq. 4.34) is so successful in
describing the kinetics of oxyanion adsorption 20,24,52 may derive from its
having, its basis in a continuous distribution of activation energies."

Table 4.2. Infrared absorption band centers for goethite, phosphated goethites, phosphate minerals, and
aqueous phosphate anions 61 ,6Z

Species
Goethite
Phosphated (H 3P04)
goethite (dry)
Phosphated (H 3P04)
goethite (wet)
Phosphated (Ca(HZP04)z)
goethite (dry)
Phosphated (Ca(HZP04)z)
goethite (wet)
(

Band centers, em"!

Species

Band centers, cm"

805, 900, 1000,


1110, 1190
734, 891, 996,
1030, 1191
1000, 1115

Strengite
(FeP0 4 2HzO)
Metastrengite
(FeP04 2HzO)
HZP04(aq)

1000, 1030, 1100,


1191
1050, 1090

HPO~-(aq)

750, 994, 1012,


1040
752, 992, 1007,
1044, 1104
878, 947, 1072,
1150, 1230
862, 988, 1076,
1230
1004

PO~-(aq)

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

143

4.5. ORGANIC MATTER ADSORPTION

The complexes formed between the solid phases in soil and native organic
compounds in the soil solution are not yet well understood because the
~

(4.45)

Table 4.3. Mechanisms of adsorption for organic compounds in soil solutions63


Mechanism
Cation exchange
......Protonation
.....Anion exchange
..... Water bridging
v Cation bridging
..... Ligand exchange
.... Hydrogen bonding
- Van der Waals interactions

Principal organic functional groups involved


Amines, ring NH, heterocyclic N
Amines, heterocyclic N, carbonyl, carboxylate
Carboxylate
Amino, carboxylate, carbonyl, alcoholic OH
Carboxylate, amines, carbonyl, alcoholic OH
Carboxylate
Amines, carbonyl, carboxyl, phenylhydroxyl
Uncharged, nonpolar organic functional groups

144

THE SURFACE CHEMISTRY OF SOilS

where B is typically a molecular unit comprising a quaternized nitrogen


atom in an aliphatic or a heterocylic aromatic structure and M+ is a
monovalent metal cation in a surface complex depicted by ==. The most
important specific example of Bin Eq. 4.45 is a proteinaceous fragment or
a carbohydrate unit containing the protonated amino group, NHt. The
metal cation M+ may be complexed by anyone of the surface functional
groups described in Sec. 1.2 under those conditions that permit either
inner- or outer-sphere complexes to form with positive ions. An important
feature of the exchange reaction in Eq. 4.45 is the subsequent "demixing"
of the two adsorbates, M+== and B+==, which is commonly observed when
the adsorbent is a 2: 1 phyllosilicate.P" Perhaps because of significant
alterations induced in the structure of adsorbed water and differences in
the stereochemistry of the inorganic cation and organocation surface complexes, the coexistence of the two kinds of complex on the same patch of
siloxane surface is not favored. Instead, a random interstratification of unit
layers, in each of which one kind of complex predominates, develops as the
exchange of M+ for B + takes place. This segregation of the two absorbates
helps to limit the extent of conformational changes the organocation
must undergo upon adsorption. If M+ in Eq. 4.45 is replaced by a bivalent
metal cation, demixing is usually reduced sharply, evidently because
organocation adsorption is now restricted to the external surfaces of
quasicrystals.
The protonation mechanism in Table 4.3 refers to the reaction whereby
an organic functional group forms a complex with a surface proton that
either is itself in an inner- or outer-sphere surface complex or is in an acidic
water molecule solvating a metal cation in an outer-sphere surface complex
(Sec. 2.4). The solid surfaces in soils can develop Breasted acidity in a
variety of ways (cation exchange, dissociation of hydroxyls and carboxyls,
hydrolysis of solvated metal cations, and so on), and this acidity offers the
possibility that proton-selective organic functional groups on dissolved
solutes can be adsorbed through a protonation reaction. This mechanism
is expected to be most important at low pH values and/or low water contents in soils, i.e., when the Brensted acidity of the solid surfaces is
greatest. Organic solutes containing amino or carbonyl groups-notably
proteinaceous materials-s- should be most susceptible to surface protonation. 63 ,64
The anion exchange mechanism in Table 4.3 is the analog of the reaction
in Eq. 4.45, wherein the signs of the valences are reversed: B symbolizes
a carboxylate group (COO-) and M is replaced by a univalent, inorganic
anion that forms outer-sphere complexes with protonated surface hydroxyl
or amine groups. This mechanism is not observed often, possibly because
of the weakness of the surface complexes involved, but it should be
prominent in acidic soils whose clay fraction comprises primarily metal
oxides.
Another weak adsorption mechanism for either anionic or polar organic
functional groups is water bridging; which involves complexation with the

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

145

proton in a water molecule solvating an exchangeable cation:


Bb-(aq) + (HzO)pM m +==

Bb-(HzO)pMm +==

(4.46)

where B is a molecular unit containing the anionic or polar moiety and


M m + is the exchangeable cation. In Eq. 4.46, b = 0 or 1, p = 3, 4, or 6
usually, and m = 1 or 2. Water briding is expected to occur particularly
when M is a relatively hard Lewis acid since, in that case, B is less likely to
displace a solvating water molecule during surface complex formation. If
displacement of the water molecule does occur, an inner-sphere complex is
formed between B b - and M m + and the adsorption mechanism is termed
cation bridging. Clearly, whether water bridging or cation bridging takes
place during the adsorption of an anionic or polar functional group by a
surface bearing solvated exchangeable cations depends on the relative
Lewis base character of the functional group and the relative Lewis acid
character of the exchangeable cation. For example, the carboxylate groups
in humic substances may adsorb on montmorillonite through cation
bridging when monovalent exchangeable cations are present and through
water bridging when bivalent exchangeable cations are present. 65
Ligand exchange refers specifically to inner-sphere complex formation
between a carboxylate group and either Al(III) or Fe(III) ,in a soil miI)~.r~1
bearing inorganic.hydroxyl groups. This mechanism is exactly analogous to
that discussed in Sec. 4.4 for inorganic oxyanion adsorption. The chemical
bonds involved are much stronger. 01. course, ~!L.that in aniq~ .S:2'change
or in the two bridging mechanisms. Evidence for the ligand exchange
mechanism in carboxylate adsorption from soil solutions is abundant but
indirect.P" The inorganic surfaces involved are 'on metal oxides and the
edges of phyllosilicates in soils.
Hydrogen bonding between organic functional groups and either siloxane oxygen atoms or surface hydroxyl groups does not appear to be a
significant factor in the adsorption of organic matter, except possibly where
it is stabilized by charge localization on the adsorbing surface (Sec. 1.2).63
On the other hand, there is evidence that this mechanism is important in
the adsorption of organic compounds by organic surfaces.P?
Of much more significance is the role of van der Waals' dispersion
interactions, which are multipole (principally dipole-dipole) iateractions
produced by correlations between fluctuating induced multipole (principally dipole) moments in two nearby uncharged, nonpolar molecules.
Although the time-averaged induced multipole moment in each molecule
is zero (otherwise it would not be nonpolar), the correlations between the
two induced moments do not average to zero, with the result that a net
attractive interaction between the two is produced at very small intermolecular distances. The van der Waals dispersion interaction between
two molecules is necessarily very weak, but when many molecules in a
polymeric structure interact simultaneously. the van der Waals component
is additive. These characteristic features of the van der Waals interaction
are discussed further in Sec. 6.2.

146

THE SURFACE CHEMISTRY OF SOILS

For polymeric organic solutes in soil solutions, the van der Waals
interaction with the atoms in a solid surface can be quite strong and
relatively long-range. The influence of this interaction in biopolymers, such
as proteins and carbohydrates, is believed to be the fundamental reason for
the very frequent appearance of Hstype adsorption isotherms when these
large molecules react with soil minerals.P" The effects of van der Waals
interactions are especially apparent when the ionic strength of the soil
solution is high enough to suppress the ionization of acidic functional
groups on large organic solute molecules or when the pH has been adjusted
to make the net charge on them vanish. (See Sec. 6.3.)
The adsorption mechanisms listed in Table 4.3 are expected to operate
when dissolved soil organic matter reacts with solid soil particles. Although
the structural chemistry of soluble organic matter in soils is not well
understood, certain generalizations concerning adsorption can be made on
the basis of studies in which the pH and surface characteristics of model
adsorbents have been varied systematically. One of the most important of
t!tese generalizations is that the quantity of dissolved organic matter
adsorbed tends to decrease as the pH increases above 4.0. 69 This fact
suggests that dissolved soil organic matter forms ligand-like surface
complexes, as described in Sec. 4.4 f~r inorganic ox~anions, and therefore
tJiat the Eredominant adsoq~tion mechanisms are those appropriate to
anions.
- The surfaces bearing charged siloxane ditrigonal cavities in soils reacts
with dissolved organic matter in two principal ways. First, the permanent,
negative surface charge produces a negative adsorption of the organic
matter that should become more pronounced as the pH value of the soil
solution increases and the organic matter becomes more anionic. Second,
carboxylate and phenolic hydroxyl groups in the organic matter particularly should form complexes with the siloxane ditrigonal cavities through
exchangeable cations in surface complexes with these cavities. These two
kinds of interaction oppose one another on the external surfaces of
quasicrystals or other aggregate units; no interlayer adsorption, negative or
positive, is involved. !!. the exchangeable cation is monovalent. the amoullt
of organic matter adsorbed increases as the Lewis acidsc;>ftness of tlJe
cation increases, indicating that the organic ligands involved are softer
Lewis bases than solvation wilter molecules and that cation bridging is
the main adsorption mechanism.F If the exchangeable cation is bivalent,
the amount of organic matter adsorbed increases as the ionic potential
of the cation increases, suggesting that weak protonation of the organic
ligands through water bridging is the principal adsorption mechanism,
since purely electrostatic interactions across a solvation shell should be
favored with cations of high ionic potential. 65 Cation bridging does not
seem likely in this case because there is no correlation between the amount
adsorbed and the Lewis acid softness of the exchangeable action.
The edge surfaces of soil phyllosilicates and the surfaces of metal
oxyhydroxides react with dissolved organic matter through the ligand

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOilS

147

exchange mechanism.?" which operates in much the same way as described


for inorganic oxyanions in Sec. 4.4. Because of the polymeric, polyfunctional nature of dissolved soil organic matter, however, it is expected that
protonation and van der Waals interactions play a role in adsorption, as
does negative adsorption when the net oxidic surface charge is negative.
Increasing soil pH both enhances negative adsorption and diminishes the
number of protons available to create the surface OH; groups that
mediate the ligand exchange reaction. Thus the acid dissociation constants
of both the organic matter and the soil adsorbents are pertinent.
Ligand exchange reactions between inorganic surface hydroxyl groups
and carboxyl groups in dissolved soil organic matter produce inner-sphere
complexes between the metal cations AI(III) and Fe(III) and the carboxyls. These reactions are relatively fast, and their effect on the metal
cation is to weaken its bonds to the oxygen ions surrounding,lt .io!he
adsorbent throu~h an accompanying redistribution of cbargr:.~ This effect
is the first step in the subsequent d,etachment of the metal-organic hgana
complex into the soil solution, the rate-determining step in the ultimate
dissolution of the solid adsorbent. In this way, the adsorption of organic
matter plays a critical role in the weathering of kaolinite, amorphous
aluminosilicates, and aluminum or iron oxyhydroxides in soils. 71
NOTES
1. The concepts and terminology of adsorption phenomena are discussed in detail
in D. H. Everett, Manual of Symbols and Terminology for Physicochemical
Quantities and Units. Appendix II: Definitions, Terminology and Symbols in
Colloid and Surface Chemistry. Butterworths, London, 1972.
2. For a brief review of experimental methods, see the first four sections in
S. Burchill, M.H.B. Hayes, and D. J. Greenland, Adsorption, in The Chemistry of Soil Processes (D. J. Greenland and M.H.B. Hayes, eds.), Wiley,
Chichester, U.K., 1978.
3. A complete discussion of the relative surface excess is given in Chap. II of
R. Defay and I. Prigogine, Surface Tension and Adsorption. Wiley, New York.
1966.
4. S. D. Forrester and C. H. Giles, From manure heaps to monolayers: One
hundred years of solute-solvent adsorption isotherm studies, Chemistry and
Industry, April 15, 1972, p. 318.
5. C. H. Giles, T. H. MacEwan, S. N. Nakhwa, and D. Smith, Studies in
adsorption. Part XI: A system of classification of solution adsorption isotherms and its use in diagnosis of adsorption mechanisms and in measurement of specific surface areas of solids, J. Chem. Soc., London, 3973 (1960).
C. H. Giles, D. Smith, and A. Huitson, A general treatment and classification
of the solute adsorption isotherm. I: Theoretical, J. Colloid Interface Sci.
47: 755 (1974). C. H. Giles, A. P. D'Silva, and I. A. Easton, A general treatment and classification of the solute adsorption isotherm. Part II: Experimental interpretation. J. Colloid Interface Sci. 47: 766 (1974).
6. The development of these two equations to describe adsorption from a~ueous
solution hall been reviewed in S. D. Forrester and C. H. Giles, op. cit. For a

148

THE SURFACE CHEMISTRY OF SOILS

comprehensive review of adsorption isotherm equations in soil chemical


studies, see C. C. Travis and E. L. Etnier, A survey of sorption relationships
for reactive solutes in soil, J. Environ. Qual. 10:8 (1981).
7. The advantages of Eq, 4.9 over other linear forms of the Langmuir equation
are discussed in J. A. Veith and G. Sposito, On the use of the Langmuir
equation in the interpretation of "adsorption" phenomena. Soil Sci. Soc. Am.
J. 41:697 (1977).
8. I.C.R Holford, RW.M. Wedderburn, and G.E.G. Mattingly, A Langmuir
two-surface equation as a model for phosphate adsorption by soils. J. Soil Sci.
25:242 (1974).
9. G. Sposito, On the use of the Langmuir equation in the interpretation of
"adsorption" phenomena. II: The "two-surface" Langmuir equation, Soil Sci.
Soc. Am. J. 46:1147 (1982). See also I. M. Klotz, Numbers of receptor sites
from Scatchard graphs: Facts and fantasies. Science 217: 1247 (1982). (In the
literature of macromolecular chemistry, Fig. 4.2 is known as a Seatchard
plot.)
10. G. Sposito, Derivation of the Freundlich equation for ion exhange reactions in
soils. Soil Sci. Soc. Am. J. 44:652 (1980).
11. R Sips, On the structure of a catalyst surface. J. Chem. Phys. 16:490 (1948).
12. See, e.g., Chap. 6 in E. Butkov, Mathematical Physics. Addison-Wesley,
Reading, Mass., 1968.
13. A thoughtful review of the problems discussed in this section is given in
R B. Corey, Adsorption vs precipitation, in Adsorption of Inorganics at
Solid-Liquid Interfaces (M. A. Anderson and A. J. Rubin, eds.). Ann
Arbor Science, Ann Arbor, Mich., 1981.
14. Demonstrations have been given in J. A. Veith and G. Sposito, op. cit.,?
G. Sposito, op. cit. ,9 and A. M. Elprince and G. Sposito, Thermodynamic
derivation of equations of the Langmuir type for ion equilibria in soils, Soil Sci.
Soc. Am. J. 45:277 (1981).
15. For a discussion of lAP, see Chap. 3 in G. Sposito, The Thermodynamics of
Soil Solutions. Clarendon Press, Oxford, 1981.
16. See, e.g., Chap. 5 in W. Stumm and J. J. Morgan, Aquatic Chemistry. Wiley,
r:">.
New York, 1981.
17. See, e.g., pp. 67ft in G. ~, op. cit.1s
18. J. W. Stucki and W. L. Banwart, Advanced Chemical Methods for Soil and
Clay Minerals Research. Reidel, Dordrecht, Holland, 1980. J. J. Fripiat,
Advanced Techniques for Clay Mineral Analysis. Elsevier, Amsterdam, 1982.
19. A review of the early sorption rate literature for phosphate is given in
J. C. Ryden and P. F. Pratt, Phosphorus removal from wastewater applied to
land, Hilgardia 48:1 (1980). See also J. A. Veith and G. Sposito.P
20. W. H. van Riemsdijk and F.A.M. de Haan, Reaction of orthophosphate with
a sandy soil at constant supersaturation, Soil Sci. Soc. Am. J. 45:261 (1981).
21. J. A. Veith and G. Sposito, Reactions of aluminosilicates, aluminum hydrous
oxides, and aluminum oxide with o-phosphate: The formation of X-ray and
amorphous analogs of variscite and montebrasite, Soil Sci. Soc. Am. J 41:870
(1977).
22. V. E. Berkheiser, J. J. Street, P.S.C. Rao, and T. L. Yuan, Partitioning of
inorganic orthophosphate in soil-water systems, CRe Critical Reviews in
Environmental Control 10:179 (1980).
23. C. C. Travis and E. L. Etnier, op. cit. 6

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

149

24. J. de Kanel and J. W. Morse, The chemistry of orthophosphate uptake from


seawater onto calcite and aragonite, Geochim. Cosmochim. Acta 42:1335
(1978). S. H. Chien and W. R. Clayton, Application of the Elovich equation
to the kinetics of phosphate release and sorption in soils, Soil Sci. Soc. Am. J.
44:265 (1980). W. H. van Riemsdijk and F.A.M. deHaan, op. cit.2o
25. C. Aharoni and M. Ungarish, Kinetics of activated chemisorption. 2: Theoretical models, J.C.S. Faraday 173:456 (1977).
26. Y.-S.R. Chen, J. N. Butler, and W. Stumm, Kinetic study of phosphate
reaction with aluminum oxide and kaolinite, Env. Sci. Technol. 7: 327 (1973).
D. N. Munns and R. L. Fox, The slow reaction which continues after
phosphate adsorption: Kinetics and equilibrium in some tropical soils, Soil Sci.
Soc. Am. J. 40:46 (1976).
27. See, e.g., Chap. 4 in W. P. Kelley, Cation Exchange in Soils. Reinhold, New
York, 1948.
28. M.G.M. Bruggenwert and A. Kamphorst, Survey of experimental information
on cation exchange in soil systems, in Soil Chemistry. B. Physico-Chemical
Models (G. H. Bolt, ed.). Elsevier, Amsterdam, 1979. M. M. Reddy, Ionexchange materials in natural water systems, Ion Exchange and Solvent
Extraction 7: 165 (1977).
29. D. G. Kinniburgh and M. L. Jackson, Cation adsorption by hydrous metal
oxides and clay, in M. A. Anderson and A. J. Rubin, op. cit. 13
30. See, e.g., Chap. 3 and 5 in G. Sposito, op. citY
31. See Sec. 6.4 in G. Sposito, op. cit. 15
32. The concept of inner-sphere surface complexation as the basis for relative
metal adsorption selectivity in soils was introduced in W. R. Heald, M. H.
Frere, and C. T. deWit, Ion adsorption on charged surfaces, Soil Sci. Soc. Am.
J. 28: 622 (1964). Further quantitative development of tlris concept was given in
I. Shainberg and W. D. Kemper, Ion exchange equilibria on montmorillonite,
Soil Sci. 103:4 (1967). See also the discussion of this paper in Soil Sci. 104:444
(1967).
33. The HSAB principle is discussed in an introductory fashion in P. J. Sullivan,
The principle of hard and soft acids and bases as applied to exchangeable
cation selectivity in soils, Soil Sci. 124:117 (1977). See also Sec. 3.3 in
G. Sposito, op. cit.15
34. M. Misono, E. Ochai, Y. Saito, and Y. Yoneda, A new dual parameter scale
for the strength of Lewis acids and bases with the evaluation of their softness,
J. Inorg. Nucl. Chem. 29: 2685 (1967).
35. A. Maes and A. Cremers, Charge density effects in ion exchange. Part 2.
Homovalent exchange equilibria, J.C.S. Faraday I 74: 1234 (1978).
36. M.G.M. Bruggenwert and A. Kamphorst, op. cit.28
37. The difficulties of this kind that arise when chloride ions are present in the
aqueous solution phase are investigated in G. Sposito, K. M. Holtzclaw,
C. Jouany, L. Charlet, and A. L. Page, Sodium-calcium and sodiummagnesium exchange on Wyoming bentonite in perchlorate and chloride
background ionic media, Soil Sci. Soc. Am. J. 47: 51 (1983).
38. A similar organization of ligand effects is presented in M. M. Benjamin and
J. O. Leckie, Conceptual model for metal-ligand-surface interactions during
adsorption, Environ. Sci. Technol. 15:1050 (1981).
39. H. Farrah and W. Pickering, The sorption of copper species by clays. I:
Kaolinite, AWl. J. Chern. 29: 1167 (1976). II: Illite and montmorillonite, AWl.

150

THE SURFACE CHEMISTRY OF SOILS

J. Chem. 29:1177 (1976). J. A. Davis and J. 0. Leckie, Effect of adsorbed

complexing ligands on trace metal uptake by hydrous oxides, Environ. Sci.


Technol. 12: 1309 (1978). See also the discussion of the third paper in Environ.
Sci. Techol. 13:1289 (1979).
40. M.D.A. Bolland, A. M. Posner, and J. P. Quirk, Zinc adsorption by goethite
in the absence and presence of phosphate, Aust. J. Soil Res. 15: 279 (1977).
M. M. Benjamin and N. S. Bloom, Effects of strong binding of anionic
adsorbates on adsorption of trace metals on amorphous iron oxyhydroxide, in
Adsorption from Aqueous Solutions (P. H. Tewari, ed.). Plenum, New York,
1981.
41. R. M. McKenzie, The adsorption of lead and other heavy metals on oxides of
manganese and iron, Aust. J. Soil Res. 18:61 (1980). H. Kerndorf and
M. Schnitzer, Sorption of metals on humic acid, Geochim. Cosmochim. Acta
44:1701 (1980). D. G. Kinniburgh, M. L. Jackson, and J. K. Syers, Adsorption of alkaline earth, transition, and heavy metal cations by hydrous oxide gels
of iron and aluminium, Soil Sci. Soc. Am. J. 40:796 (1976). H. Farrah and
W. F. Pickering, Influence of clay-solute interactions on aqueous heavy metal
ion levels, Water, Air and Soil Pollution 8: 189 (1977).
42. D. G. Kinniburgh and M. L. Jackson, op. cit.29 H. Farrah and W. F. Pickering, op. cit.39 D. G. Kinniburgh and M. L. Jackson, Concentration and pH
dependence of calcium and zinc adsorption by iron hydrous oxide gel, Soil Sci.
Soc. Am. J. 46: 56 (1982).
43. M. H. Kurbatov, G. B. Wood, and J. D. Kurbatov. Isothermal adsorption of
cobalt from dilute solutions, J. Phys. Chem. 55:1170 (1951). The term was
coined in D. G. Kinniburgh and M. L. Jackson, op. cit. ,29 p. lOI.
44. M. M. Benjamin and J. 0. Leckie, Effects of complexation by CI, S04, and
S203 on adsorption behavior of Cd on oxide surfaces, Environ. Sci. Technol.
16: 162 (1982).
45. J.T.G. Overbeek, Electrokinetic phenomena, in Colloid Science, Vol. I
(H. R. Kruyt, ed.). Elsevier, Amsterdam, 1952. R.O. James and
T. W. Healy, Adsorption of hydrolyzable metal ions at the oxide-water
interface. II: Charge reversal of Si02 and Ti0 2 colloids by adsorbed Co(I1),
La(I1I) , and Th(IV) as model systems, J. Colloid Interface Science 40: 53
(1972). c.-P. Huang and W. Stumm, Specific adsorption of cations on hydrous
a-Ah03, J. Colloid Interface Science 43: 409 (1973). S. L. Swartzen-Allen and
E. Matijevic, Colloid and surface properties of clay suspensions. II: Electrophoresis 'and cation adsorption of montmorillonite, J. Colloid Interface Sci.
50:143(1975). G. R. Wiese, R. 0. James, D. E. Yates, and T. W. Healy,
Electrochemistry of the colloid-water interface, in Electrochemistry (J. O'M.
Bockris, ed.). Butterworths, London, 1976. D. W. Fuerstenau, D. Manmohan, and S. Raghavan, The adsorption of alkaline-earth metal ions at the
rutile/aqueous solution interface, in P. H. Tewari, op. cit.40
46. D. W. Fuerstenau et al., op. cit.45
47. R. 0. James and T. W. Healy, op. cit. 45
48. See, e.g., M. M. Benjamin and J. 0. Leckie, Adsorption of metals at oxide
interfaces: Effects of the concentrations of adsorbate and competing metals, in
Contaminants and Sediments, Vol. 2 (R. A. Baker, ed.), Ann Arbor Science,
Ann Arbor, Mich., 1980.
49. See, e.g., p. 51 in F. C. Loughnan, Chemical Weathering of the Silicate
Mtnerals. American Elsevier. New York. 1969.

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS

151

50. C.J.B. Mott, Anion and ligand exchange, in D. J. Greenland and M.H.B.

Hayes, op, cit.2


51. R. L. Parfitt, Anion adsorption by soils and soil materials, Advan. Agron. 30: 1
(1978).
52. F. J. Hingston, A review of anion adsorption, in M. A. Anderson and A. J.

Rubin, op. cit. 13


53. R. E. White, Retention and release of phosphate by soil and soil constituents,
in Soils and Agriculture (P. B. Tinker, ed.). Wiley, New York, 1981.
54. J. A. Davis and J. O. Leckie, Surface ionization and complexation at the
oxide/water interface. 3: Adsorption of anions, J. Colloid Interface Sci. 74: 32
(1980).
55. See, e.g., J. R. Sims and F. T. Bingham, Retention of boron by layer silicates,
sesquioxides, and soil materials: I: Layer silicates, Soil Sci. Soc. Am. J. 31: 728
(1967). III: Iron- and aluminum-coated layer silicates and soil materials, Soil
Sci. Soc. Am. J. 32: 369 (1968).
56. J. R. Sims and F. T. Bingham, Retention of boron by layer silicates, sesquioxides, and soil materials, Soil Sci. Soc. Am. J. 32: 364 (1968). M. A. Anderson,

57.

58.

59.

60.
61.

62.

J. F. Ferguson, and J. Gravis, Arsenate adsorption on amorphous aluminum


hydroxide, J. Colloid Interface Sci. 54: 391 (1976). F. J. Hingston, A. M.
Posner, and J. P. Quirk, Competitive adsorption of negatively charged ligands
on oxide surfaces, Disc. Faraday Soc. 52: 334 (1971).
D. T. Malotky and M. A. Anderson, The adsorption of the potential determining arsenate anion on oxide surfaces, Colloid Interface Sci. 4: 281 (1976).
M. A. Anderson and D. T. Malotky, The adsorption of protolyzable anions on
hydrous oxides at the isoelectric pH, J. Colloid Interface Sci. 72: 413 (1979).
R. J. Atkinson, A. M. Posner, and J. P. Quirk, Kinetics of isotopic exchange
of phosphate at the a-FeOOH-aqueous solution interface, J. Inorg. Nucl.
Chem. 34: 2201 (1972). J. H. Kyle, A. M. Posner, and J. P. Quirk, Kinetics of
isotopic exchange of phosphate adsorbed on gibbsite, J. Soil Sci. 26: 32 (1975).
Note that the parameters A and B in these two papers correspond to k l and k 2
in Eq. 4.34.
A comprehensive discussion of this and other aspects of the ligand exchange
reaction in dissolution-precipitation reactions is given in W. Stumm, G. Furrer,
and B. Kunz, The role of surface coordination in precipitation and dissolution
of mineral phases, Croatica Chem. Acta 58:593 (1983).
D. E. Yates and T. W. Healy, Mechanism of anion adsorption at the ferric and
chromic oxide/water interfaces, J. Colloid Interface Sci. 52: 222 (1975).
R. L. Parfitt, J. D. Russell, and V. C. Farmer, Confirmation of the surface
structures of goethite (a-FeOOH) and phosphated goethite by infrared
spectroscopy, i.c.s. Faraday I 72: 1082 (1976). R. L. Parfitt, Phosphate
adsorption on an oxisol, Soil Sci. Soc. Am. J. 41: 1065 (1977). R. L. Parfitt,
R. J. Atkinson, and R. St. C. Smart, The mechanism of phosphate fixation on
iron oxides, Soil Sci. Soc. Am. J. 39: 837 (1975). R. L. Parfitt, The nature of the
phosphate-goethite (a-FeOOH) complex formed with Ca(H2P04h at different
surface coverage, Soil Sci. Soc. Am. J. 43:623 (1979). J. B. Harrison and V. E.
Berkheiser, Anion interactions with freshly prepared hydrous iron oxides,
Clays and Clay Minerals 30: 97 (1982).
The data in Table 4.2 are extracted in part from a compilation in S. R.
Goldberg, A Chemical Model of Phosphate Adsorption on Oxide Minerals and
Soils. Ph.D. dissertation. University of Califcrnia, Riverside. 1983.

152

THE SURFACE CHEMISTRY OF SOilS

63. A summary of the earlier studies with model compounds is in the classic review
by M. M. Mortland, Clay-organic complexes and interactions, Advan. Agron.
22: 75 (1970). See also D. J. Greenland, Interactions between humic and fulvic
acids and clays, Soil Sci. 111:34 (1971).
64. See, e.g., Chap. 7 in B.K.G. Theng, Formation and Properties of ClayPolymer Complexes. Elsevier, Amsterdam, 1979.
65. See Chap. 12 of B.K.G. Theng, op. cit. ,64 and Clay-polymer interactions:
Summary and perspectives, Clays and Clay Minerals 30: 1 (1982).
66. R. L. Parfitt, A. R. Fraser, J. D. Russell, and V. C. Farmer, Adsorption on
hydrous oxides. II: Oxalate, benzoate and phosphate on gibbsite, J. Soil Sci.
28:40 (1977). R. L. Parfitt, A. R. Fraser, and V. C. Farmer, Adsorption on
hydrous oxides. III. Fulvic acid and humic acid on goethite, gibbsite, and
imogolite, J. Soil Sci. 28: 289 (1977). R. Kummert and W. Stumm, The surface
complexation of organic acids on hydrous a-AI2 0 3 , J. Colloid. Interface Sci.
75: 373 (1980). S. N. Yap, R. K. Mishra, S. Raghavan, and D. W. Fuerstenau,
The adsorpton of oleate from aqueous solution onto hematite, in P. H. Tewari,
op. cit/" J. A. Davis, Adsorption of natural dissolved organic matter at the
oxide/water interface, Geochim. Cosmochim. Acta 46: 2381 (1982).
67. See, e.g., S. Burchill et al., op. cit.2 pp. 325ff.
68. See, e.g., Chaps. 7 and 10 of B.K.G. Theng, op. cit.64
69. Chapter 12 in B.K.G. Theng, op. cit.64 J. A. Davis, op. cit.66 J. A. Davis, in
Vol. 2 of R. A. Baker, op. cit.48
70. P. Chassin, N. Nakaya, and B. Le Berre, Influence des substances hurniques
sur les proprietes des argiles. II. Adsorption des acides humiques et fulviques
par la montmorillonite, Clay Minerals 12: 261 (1977). R. L. Parfitt et aI.,
op. cit.66 K. R. Tate and B.K.G. Theng, Organic matter and its interactions
with inorganic soil constituents, in Soils with Variable Charge (B.K.G. Theng,
ed.) New Zealand Society of Soil Science, Lower Hutt, N.Z., 1980.
71. K. H. Tan, The catalytic decomposition of clay minerals by complex reaction
with humic and fulvic acid, Soil Sci. 120: 188 (1975). K. R. Tate and B.K.G.
. 70
Th eng, op. CIt.

FOR FURTHER READING

M. G. Browman and G. Chesters, The solid-water interface: Transfer of organic


pollutants across the solid-water interface. Advan. Environ. Sci. Techno!. 8(1): 49
(1977). The adsorption mechanisms for pesticides in soils are discussed in the first
three sections of this review article.
C. H. Giles and S. D. Forrester, Studies in the early history of surface chemistry,
Chemistry and Industry, Nov. 8, 1969; Jan. 17, 1970; Jan. 9, 1971; July 24, 1971;
Nov. 13,1971; April 15,1972. This history of surface phenomena is a must for any
serious student of the subject. Parts V and VI, on solute adsorption by solid
surfaces, give and well-illustrated, lively account of the role of soil chemists in the
development of the adsorption isotherm concept.
D. J. Greenland and M.H.B. Hayes, The Chemistry of Soil Processes. Wiley,
Chicester, U.K., 1981. Chapters 5 and 6 in this comprehensive treatise provide
detailed surveys of experimental methods and molecular mechanisms for adsorption phenomena in soils.

INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOilS

153

M. M. Mortland, Clay-organic complexes and interactions, Advan. Agron. 22: 75


(1970). This classic review article remains the best comprehensive introduction to
adsorption mechanisms for organic compounds.
B.K.G. Theng, Formation and Properties of Clay-Polymer Complexes. Elsevier,
Amsterdam, 1979. Chapters 7, 10, and 12 of this encyclopedic treatise present
detailed discussions of the adsorption of compounds in soil organic matter by clay
minerals.

5
CHEMICAL MODELS OF
SURFACE COMPLEXATION

5.1. THE DIFFUSE DOUBLE LAYER MODEL

Diffuse double layer (DDL) theory as applied to the surface chemistry of


soils refers to the description of ion charge and inner potential contained in
the Poisson-Boltzmann equation.'
(5.1)
where /I(x) is the inner potential at a distance x from the surface of a soil
particle along a normal extending into the soil solution. The sum includes
all charged species in the soil solution, with species i having the valence
Z, and the bulk concentration c, in moles per cubic meter. The other
parameters are as described following Eq. 1.11.
Besides the hypothesis that the Poisson equation (Eq. 3.26), from which
Eq. 5.1 is derived, is physically meaningful when x is measured over
molecular dimensions, there are four basic assumptions embodied in the
Poisson-Boltzmann equation as written above:
1. The surface from which x is measured is a uniform, infinite plane of
charge characterized by a density a expressed in coulombs per square
meter.
2. The charged species in the soil solution are point ions dissociated
completely from the planar surface lying at x = O. These ions interact
among themselves and with the surface through the coulomb force.
3. The water in the soil solution is a uniform continuum liquid characterized by the dielectric constant, D.
4. The inner potential, /I(x), is proportional to W/(x), the average energy
required to bring an ion i from a point at infinity to a point at x in
the soil solution. Since W,(x) should include the effects of both noncoulombic interactions (e.g., short-range repulsive forces that deter-

CHEMICAL MODELS OF SURFACE COMPLEXATION

155

mine ion size) and fluctuations of the true inner potential about its mean
value, l/J(x), it follows that these effects are neglected when the DDL
assumption
(5.2)

is invoked.
The limitations imposed on DDL theory as a molecular model by these
four basic assumptions have been discussed frequently and remain the
subject of current research.v? In Sees, 1.4 and 3.4 it is shown that DDL
theory provides a useful framework in which to interpret negative adsorption and electrokinetic experiments on soil clay particles. This fact suggests
that the several differences between DDL theory and an exact statistical
mechanical description of the behavior of ion swarms near soil particle
surfaces must compensate one another in some way, at least in certain
applications. Evidence supporting this conclusion is considered at the end
of the present section, whose principal objective is to trace out the broad
implications of Eq. 5.1 as a theory of the interfacial region. The approach
taken serves to develop an appreciation of the limitations of DDL theory
that emerge from the mathematical structure of the Poisson-Boltzmann
equation and from the requirement that its solutions be self-consistent in
their physical interpretation. The limitations of DDL theory presented in
this way lead naturally to the concept of surface complexation.
The surface charge density that accumulates in a plane lying at a distance
x from a soil particle surface along a normal extending into the soil solution
can be calculated with the equation!
u(x)

= (00 ~ CiZiF exp( - ZiFl/J(X')jRT)dx'

Jx

(5.3)

In Eq. 5.3, it is assumed that the concentration of each ion in the soil
solution achieves its "bulk value", c., well before another soil particle
surface is encountered along the x direction moving out from the origin of
spatial coordinates at x = O. Thus the distance separating soil particle
surfaces is assumed to be much larger than the domain over which l/J(x)
differs significantly from zero. 3 With this assumption, the upper limit of the
integral defining u(x) can be set at infinity and the boundary condition,
that the DDL potential and electric field intensity vanish at this upper
limit, can be applied.
The mathematical identity
2l/J

d(dl/J) == d 2 dx
dx
dx

(5.4a)

and Eq, 5.3 can be used to convert Eq. 5,1 to the differential equation
dl/J

u(x)

dx

enD

-=--

(5,5)

156

THE SURFACE CHEMISTRY OF SOILS

where the vanishing of the electric field intensity, -dl/J/dx, at a point at


infinity has been invoked. This formal result can be developed into an
explicit equation for O"(x) by integrating both sides of the PoissonBoltzmann equation with the help of another mathematical identity:
d2l/J
dl/J) 2
d ( dx == 2 dx2 dl/J

(5.4b)

The integration of Eq. 5.1 with respect to l/J from l/J((0) = 0 to l/J(x) yields
the differential equation
2

dl/J) = -2RT L dexp( -ZiFl/J(X)/RT) - 1]


( -d
x
eoD i

(5.6)

The square root of both sides of Eq. 5.6 can be taken under the convention
that the sign of the root is opposite to that of l/J(x). Then Eqs. 5.5 and 5.6
produce the expression
1/ 2

O"(x) = -sgn(l/J) { 2eoDRT ~ dexp( -ZiFl/J(X)/RT) - 1]}

(5.7)

where
sgn( l/J)

+1

= { -1

l/J>O
l/J<O

Equation 5.7 is a generalization of Eq. 3.39, which was used in the


interpretation of electrokinetic phenomena. It establishes the DDL model
relationship between the electric potential at a point and the accumulated
density of surface charge at the point, subject to the condition that 0" vanish
with l/J.
Equation 5.5, a general expression that relates the electric field to the
density of surface charge in a system exhibiting rectangular symmetry, 1 is
equivalent to the differential equation
2
dl/J
{2RT
(5.8)
dx = -sgn(l/J) eoD~ ci[exp(-ZiFl/J(X)/RT) -1]

}1/

according to Eq. 5.7. The solution of this equation to obtain l/J(x) permits
the calculation of O"(x) and the volumetric charge density,
p(x) =

CiZiF exp( -ZiFl/J(X)/RT)

(5.9)

These two charge densities and the inner potential provide a complete
description of the interfacial region according to DDL theory.
Equation 5.8 can be solved analytically in three special cases wherein the
aqueous solution phase contains a single strong electrolyte." These are the
symmetric electrolyte (e.g., NaCI0 4), the 2: 1 electrolyte, (e.g.,
Ca(CI04h ), and the 1: 2 electrolyte (e.g., Na2S04)' The mathematical
manipulations involved are simplified after the transformations

y Fl/J/RT

f3 2F 2/ F. oDRT

157

CHEMICAL MODELS OF SURFACE COMPLEXATION

introduced in connection with Eq. 1.17, are applied to Eq. 5.8:

dy

dx = -sgn(y)J{i{c+[exp(-Z+y) - 1] + c[exp(-Z_y) - I]P/2

(5.10)

where c., and Z+ refer to the cation and c: and Z_ refer to the anion in the
electrolyte. Equation 5.10 is the working DDL equation to be solved in the
three special cases. The particular forms of this equation are

Symmetric electrolyte: Z+ == Z = -Z_, C+

dy
dx

= Co = c:

-sgn(y)K[exp( -Zy) + exp(Zy) - 2r/2

= - K exp( -Zy/2) [exp (Zy) - 1]

2: 1 electrolyte: Z+ = 2, Z_ = -1, C+

= Co, c:

(5.lIa)

== 2co

dy

dx = -sgn(y)K[exp( -2y) + 2 exp(y) - 3]1/2

+ 2 exp(y)]l/2 (5. Llb)


1:2 electrolyte: Z+ = 1, Z_ = -2, c., == 2co, c: == Co
= -K exp( -y)[exp(y) - 1][1

dy

dx = -sgn(Y)K'[2 exp( -y) + exp(2y) - 3]1/2

= K' exp(y) [exp( -y) - 1][1 + 2 exp( _y)]1/2

(5.lIc)

where K = J f3co and K' = J f3co. The solutions of these equations are
listed in Table 5.1. As an illustration of the method by which they are
obtained, consider Eq. 5.lIa. A separation of variables and the subsequent
integration of both sides of this equation yields the result
Y(X) eZy/2 dy
eZY - 1 =

1 iY(X)

2:

i Yo

csch(Zy/2)dy
Yo

.!.In[tanh(ZY(X)/4)]
Z
tanh (ZYo/4)

-K

J: dx' = -KX

where
cschu=

2
U

e - e

e" - e- u
tanh u = U
-u
e +e

and Yo ;& yeO). Upon rearranging this result to be an equation for ",(x), one
obtains
",(x) = 4 -RT tanh- 1(ae- Z.'<X)
ZF

(5.12)

158

THE SURFACE CHEMISTRY OF SOILS

Table 5.1. Analytical solutions of the Poisson-Boltzmann equation for single

electrolytes''
Electrolyte

DDL electric potential"

Symmetric

l/J(X) = (4RTjZF)tanh- 1(ae-ZKX ) with a = tanh(ZFl/J(O)j4RT)

2:1

l/J(x) =

(RTjF)ln[~ tanh

(V;

KX +

b) -~]

with b = tanh- 1{[1 + 2 exp(Fl/J(O)jRT)F/ 2j J3}


2: 1

6b' exp(J3Kx) }
l/J(x) = (RTjF) In { 1 + [b' exp(J3Kx) - IF

.
(1 + 2eYO)1/2 + J3
With b' = (1 + 2eYO)1/2 _ J3' Yo = Fl/J(O)jRT

1:2

6c exp(J3K'X) }
l/J(x) = -(RTjF) In { 1 + [c exp(J3K'X) _ 1)2
.
(1 + 2e- YO)1/2 + J3
with c = (1 + 2e YO)1/2 _ J3' Yo = Fl/J(O)jRT

1:2

l/J(x)

-(RTjF)

In[~ tanh

(1

K'X +

C') -~]

with c' = tanh- I{[1 + 2 exp(-Fl/J(O)jRT)]1/2jJ3}


= J f3c+ and 1<' = J f3c-. When two entries appear for ",(x), the first applies to negative potentials,
the second to positive potentials.

I<

where a == tanh(ZFr/1(O)/4RT). The corresponding surface charge density


follows from Eq. 5.5:
8FKa exp(-ZKX)
O'(X) =
,8[1 - a2 exp( -2ZKX)]
(5.13)
where the formula

has been used. The inverse hyperbolic tangent has the MacLaurin expansion
u3
tanh-1u = u + - + ...
3
which, for large values of the distance x (i.e., ZKX
to be replaced by the approximation
RT
r/1(x) == 4 ZF a exp( - ZKX)

1), permits Eq. 5.12

(5.14)

Equation 5.14 shows that the DOL inner potential decreases exponentially

CHEMICAL MODELS OF SURFACE COMPLEXATION

159

in absolute magnitude at very large distances. Note that this decrease is


more rapid as the valence Z increases.
The solutions of Eqs. 5.11b and 5.11c are found in a similar manner."
(Note that Eq. 5.11c can be derived from Eq. 5.11b by replacingy with-y
everywhere.) For these cases, the mathematical form of the solution
depends on whether the DDL potential is negative or positive. However,
in each case it can be demonstrated that the absolute magnitude of the
potential approaches its asymptotic value exponentially, as in Eq. 5.14.
The applicability of the analytical results in Table 5.1, or of any solution
to Eq. 5.8 obtained by numerical integration, to the chemical modeling
of the interfacial region can be examined in the context of the balance
of surface charge, discussed in Sec. 3.1. The obvious relevance of DDL
theory to surface charge balance is as a method for estimating 0'0, the
equivalent surface density of dissociated charge. If it is reasonable to locate
this surface density in a single plane lying at a distance x from a soil particle
surface, then 0'0 = O'(x) and
0'0

= -sgn("') {2e oDRT

~ e;[exp( - ZiF",(X)/ RT) -

1]f/2 (5.15)

according to Eq. 5.7. The special case of Eq. 5.15 that occurs when x is the
was presented
position of the electrokinetic plane of shear (and ",(x) =
in Eq. 3.39. Equation 1.11, which figures in the DDL model of negative
anion adsorption, can be derived from Eq. 5.15 under the assumptions that
the soil solution contains only a 1: 1 electrolyte and that the diffuse ion
swarm comes into contact with a soil particle in the plane x = 8. The
explicit x dependence of 0'0 is then

-8FK

O'o(x)

= f3

a exp( -KX)
1 _ a2 exp( -2KX)

(x

= 8)

(5.16)

according to Eq. 5.13.


A detailed model of the interfacial region requires the specification of
the position of the plane where the diffuse ion swarm begins. A popular
choice in the literature of soil chemistry' has been x = 0, which means that
outer-sphere surface complexes are neglected entirely and inner-sphere
surface complexes are ignored if they would protrude beyond the plane to
which O'in' the intrinsic surface charge density, refers. (See Sees. 1.5 and
3.1 for a discussion of O'in') That this choice is not reasonable physically,
however, can be seen from a simple calculation involving Eq. 5.16.
Consider a 1: 1 electrolyte at the concentration Co = 100 mol m- 3 and
suppose that ",(0) = -8RT/F, a value that is not unrealistic for a smectite siloxane surface. Then K = J f3co = 1.04 X 109 m- 1 at 298 K, a =
tanh (-2) = -0.96403, and
.
0.96403
8(9.6487)(1.04) 10- 3 x
O'D(O) =
1.084
x
1 - (0.96403)2
- 1.01 Cm

THE SURFACE CHEMISTRY OF SOilS

160

according to Eq.5.16, with F = 9.6487 X 104 Cvmol ?


and f3 =
16
1.084 X 10 m mol"! at 298 K. This result is about 10 times greater than a
typical value of 10"01 for a smectite, according to the calculation presented in
Eq. 1.22. It is therefore impossibly large.
The root cause of the difficulty encountered when the DDL model is
applied at the surface to which O"in refers is the neglect of ion size effects. If
IO"in I were as large as what DDL theory predicts, the soil particle would be
very likely to form inner- and outer-sphere surface complexes that would
decrease lO"pl in Eq. 3.2 and lead to a choice of x > O. For example, if
the formation of outer-sphere complexes on a smectite siloxane surface required that the diffuse ion swarm in a 0.1 M 1: 1 electrolyte begin
at x = 0.7 nm, then a calculation of 0"0(0.7 nm) would result in
0"0 = 0.032 C om- 2 , a reasonable estimate of lO"pl. Looking at the computation in reverse, one could set 0"0 = 2/310"91, according to the estimate
of the effect of outer-sphere surface complexation on Na-montmorillonite
in NaC! given in connection with the discussion of Table 1.8. The
corresponding value of x in Eq. 5.16 would then be x = 0.5 nm, which is
again quite reasonable. These examples should make it clear that Eq. 5.15
must be understood to refer to a point somewhere around 0.5 nm from the
surface bearing the intrinsic charge on a soil particle in order that the DDL
model be realistically applicable to the interfacial region. 6
A more elegant way to illustrate the constraint that 0"0 in Eq. 5.15 be
evaluated at a point at least 0.5 nm from the "bare" soil particle surface
comes from the results of Monte Carlo computer simulations of ionic
solutions near charged planes." This technique, which was mentioned in
Sec. 2.1 in connection with the simulation of the I structure in liquid water
(Fig. 2.2), permits the calculation of the molecular configuration in any
physical system of known density, given its intermolecular potential
function. In the case of the DDL model, an appropriate physical system is
a set of hard-sphere cations and anions immersed in a dielectric continuum
and interacting with a charged plane and among themselves as a result of
the coulomb force." Because of the finite size of the ions, their centers
cannot be found closer to the charged plane than a distance equal to their
hard-sphere radius, and the results of a computer simulation of their
molecular behavior must be compared with the solution of Eq. 5.8
restricted to x values larger than this radius. The same condition applies to
Eqs. 5.9 and 5.15. Figure 5.1 shows the x dependence of the components
of p(x) in Eq. 5.9 as deduced from Monte Carlo computer simulations of
1 : 1 and 2: 1 electrolytes near a negatively charged plane. 7 The simulations
were made for cations and anions with a radius of 0.213 nm immersed in a
continuum with a dielectric constant equal to 78.5 (that of liquid water at
298 K). The concentration of the 1: 1 electrolyte was 100 mol- m -3, and
the ions confronted a plane whose surface charge density equaled
0.266 Com -2. The concentration of the 2: 1 electrolyte was 50 mol- m -3,
and it confronted a plane of charge density 0.177 C om- 2 For the 1: 1
electrolyte. there is excellent agreement between the data points provided
0

3~

I: I ELECTROLYTE
Co =100 mol m- 3
CT =-0.266C m- 2

3
0

o
C\J

<,

<,

Q.. 2
~

0
0

0
0

Q... 2

CATIONS

<,

CATIONS

2:1 ELECTROLYTE
c 0 =50molm- 3
CT =-0.177C m- 2

0
0

<,

Q..

10

15

x/d {d=0.425nm}

20

1
0'

ANIONS.

x/d {d=0.425nm}

Ylpre 5.1. Graphs of the component volumetric charge densities, p+(x) and p_(x), based on Monte Carlo computer simulation (data
points) and DDL theory (solid lines)."

162

THE SURFACE CHEMISTRY OF SOILS

by the computer simulation and the values of the volumetric charge density
components,
p+(x) = coexp(-FIjJ(x)jRT)
p_(x) = coexp(FIjJ(x)jRT)

(5.17)

calculated for Co = 100 mol-rn"? at 298 K with the help of Eq. 5.12 ..
At this low concentration, the errors inherent in Eq. 5.1 appear to be mutually compensating.f At higher concentrations (e.g., Co = 103 mol, m- 3 ) ,
the computer simulation results deviate significantly from the predictions
of Eq. 5.17, in that DDL theory underestimates the extent of negative
anion adsorption and fails to reproduce the oscillation in p + (x) produced
by fluctuations in the true electric potential about its mean value, ljJ(x). 7
For the 2: 1 electrolyte, these inadequacies of DDL theory are apparent
even at Co = 50 mol-rn ":', as shown in Fig. 5.1. In this case, the DDL
model predictions of ion distribution have only qualitative significance,
even when corrected for finite ion size by restricting them to the region
x > dj2, where d is the ionic diameter. The results of computer simulation
indicate that the Poisson-Boltzmann equation does not provide an accurate
description of ion swarms containing bivalent species."

5.2. SURFACE COMPLEXATION MODELS:


STATISTICAL MECHANICS
Consider a solid surface on which a single ionic species has been adsorbed
from an aqueous solution phase through the mechanism of either innersphere or outer-sphere surface complex formation. The solid surface is
assumed to bear only one kind of reactive functional group, there being M
of these groups per unit area, and it is further assumed that only one group
participates in a complex with an adsorbed molecular unit. The geometric
properties of the surface do not have to be prescribed more closely than by
this requirement of a one-to-one complexation reaction with adsorptive
ions and by the condition that it be reasonable to speak of at least a
well-defined average of z nearest-neighbor reactive functional groups for
each such group on the surface.
These simplifying assumptions concerning ionic adsorption mediated by
a surface complexation reaction make it possible to give a straightforward
foundational discussion of the principal molecular-theoretic features of
several recent surface complexation models. These models, whose specific
properties are considered in Sees. 5.3 to 5.6, have as their overall objective
the quantitative description of adsorption phenomena in terms of the
structure of the interfacial region imparted by the formation of inner- and
outer-sphere surface complexes. Just as the DDL model attempts to give
quantitative significance to the term O'D in the balance of surface charge
through a specification of the spatial distribution of surface-dissociated

".
.!
"

CHEMICAL MODELS OF SURFACE COMPLEXATION

163

ions, the surface complexation models attempt to give quantitative meaning to the terms (TIS and (T as in the balance of surface charge through
predictions of the speciation of surface-complexed ions. The combination of these two models then leads to a computer-based algorithm for the
estimation of adsorption isotherms and related surface chemical data."
The fundamental molecular attributes of the leading surface complexation models can be revealed by pursuing a statistical mechanical description
of the chemical system introduced in the first paragraph. Recall that
statistical mechanics is the branch of physical chemistry that applies the
concepts of probability to interpret the behavior of matter in stable
states. 10 For the present discussion, the relevant probability postulate from
statistical mechanics is embodied in the Gibbs factor for a single molecular
system:
P(system)

Q exp(Np,/ k B T)

(5.18)

where N is the number of molecules in the system, each of which has the
chemical potential (per molecule) u: kB = 1.38054 X 10- 23 J. K- 1 is the
Boltzmann constant; T is absolute temperature, and Q is the partition
function for the system. The partition function can be viewed as the
relative probability that the system is in any of the numerous states of
differing energy available to it, whereas the Gibbs factor, P(system) ,
represents the relative probability that the system exists with N identical
molecules in any of these states.l" Thus P(system) is interpreted as the
relative likelihood of observing a chemical system in equilibrium with a
thermal reservoir, at temperature T, that regulates the internal energy,
and with a matter reservoir, at chemical potential p" that regulates the
number of molecules. When the chemical system is an array of surface
complexes, the matter reservoir is just the contiguous aqueous solution
phase that contains the adsorptive ions.
Suppose that the surface complexes and the remaining uncomplexed
surface functional groups in the prototypical chemical system under
consideration do not interact with one another, except "geometrically"
through the constraints

= N SR + NSR'c
zNSR
2Nc c + NRC = zNSR'c

M
2NR R + NRC =

(5.19a)
(5.19b)

where SR designates an uncomplexed surface functional gro~p, SR'C


designates a surface complex, Ni(i = SR or SR'C) is the number of species
(per unit area) of type i, and Nj k = Nk j (j, k = Rand C) is the number of
pairs of nearest-neighbor surface species (per unit area) in which one
species is of type j and one is of type k, with R designating a surface
functional group and C a surface complex. Equations 5.19 represent
conditions imposed by the geometric requirements of a fixed total number
of surface functional groups M and a one-to-one surface complexation
reaction.!" Because these M groups are assumed to react independently, it

164

THE SURFACE CHEMISTRY OF SOILS

is sufficient to formulate a statistical mechanical description of them by


considering the two chemical options available to a single surface functional group. The relative probability that an uncomplexed group exists on
the surface is
(5.20a)
and the relative probability that a surface complex exists instead is
PSR'C =

QSR'C

exp(P-c/ kBT)

(5.20b)

according to Eq. 5.18. The relative probability that either an uncomplexed


group or a complex will be found on the surface at a given location is the
sum of the relative probabilities in Eq. 5.20:
{(T,P-R,P-c) =

QSR

exp(P-R/kBT) +

QSR'C

exp(p-c/kBT) (5.21)

The function {(T, P- R ,p-c) provides the basis for calculating any property
of the system comprising M independent surface functional groups.
If the surface functional group on a solid particle is electrically charged
(e.g., the siloxane ditrigonal cavity on montmorillonite), or if an innersphere surface complex formed on it bears a net charge (e.g., a phosphate
anion complexed on goethite), it is not reasonable to assume no interactions among the surface species. The same conclusion applies to any
outer-sphere surface complex as well, since the mere physical separation of
a complexed, charged species (e.g., a metal cation) from an ionized
functional group (e.g., a dissociated surface hydroxyl) leads to a nonvanishing resultant electric field, even in the case of a neutral complex. At
the very least, one expects that relatively long-range, although partially
screened, coulomb forces will operate among charged functional groups Or
complexes on a surface, just as they do in an aqueous solution phase.
These electric forces can be envisioned, in a first approximation, to be of a
nonspecific character that produces a kind of background potential field
that mediates the interactions among the surface species. Taking the
simplest case, one can estimate the strength of the background potential by
the product of the average number of like nearest neighbors of a central
surface species and the average energy of interaction between the central
species and one of its neighbors.P
Ziet/Ji == (zNJM)ei = average number of nearest neighbors x average
interaction energy per neighbor
where Z, is the species valence and t/Ji is the average electric potential
created by z nearest neighbors of the surface species i. However, the
potential field, ,pi, should not be limited to the coulomb effects produced
by nearest neighbors, since it pertains to long-range interactions. A more
appropriate expression for t/Ji can be obtained by assuming that its van der
Waals limit exists. This limit refers to the mathematical process of letting z
become infinite while F/ goes to zero in such a manner that their product.

CHEMICAL MODELS OF SURFACE COMPLEXATION

165

1jJ? = UJZie, remains finite:

(5.22)
in the van der Waals limit. Equation 5.22 represents the potential field
obtained when the range of interaction tends to infinity and the strength of
interaction tends to zero in a special way. The concept of a long-range
average potential field is not limited to coulombic interactions. It also
provides the physical basis for the well-known van der Waals models of
liquids, solutions, and ferromagnets.l"
The effect of the average long-range potential field, IjJ(VW) , on P(systern) in Eq. 5.20 is to multiply Q by the Boltzmann factor,
exp( - ZeljJ(vw) j T), where Z is the valence of the molecular species
under consideration: 10
p(VW)(system) == Q exp(-ZeljJ(vw)jkBT)exp(p,fkBT)

(5.23)

The Boltzmann factor acts to modify p(vw)(system) for the effect of the
"external potential field", ljJ(vw), which acts on each surface species. The
corresponding change in g in Eq. 5.21 is
g(vw)(T,P.R,P.c)

QSR exp[(p.R - ZSR e ljJ~v;)jkBT]


+QsR'c exp[(p.c - ZSR'C e 1jJ~~DjkBT] (5.24)

Equation 5.24 can serve as an approximate relative probability expression


for an array of interacting surface species. Note that only the long-range
part of the interactions is given explicit consideration. For example, if the
surface functional group bears no net charge (e.g., a surface hydroxyl
group), then no lateral interaction among such groups appears in the van
der Waals model since ZSR equals zero in this case.
For the binary surface chemical system described by Eq. 5.24, the most
important special cases of complexation reactions are
SRZSR(S) + pMm+(aq) + qL1-(aq) + xH+(aq)

+ yOH-(aq)

= SR'Mp(OH)yHxL~SR'C

SRZSR(S) + qL1-(aq) + xH+(aq)

+ pZp(aq)

= SHxL~sC

(5.25a)

+ Rr-(aq) (5.25b)

Several examples of these two surface complexation reactions are listed in


Table 5.2. The reaction in Eq. 5.25a describes either inner-sphere complexation of a cationic species or outer-sphere complexation of any ionic
species. Electroneutrality requires that
ZSR + pm + x - ql - y

ZSR'C + Zp

(5.26)

where m is the valence of the reacting metal M, -I is the valence of the


reacting ligand L (either inorganic or organic), and SR is assumed to
comprise SR'(s), an undissociable moiety, and P. a dissociable molecular
unit whose valence is Z,J' (The valence of SR must be equal to
I

Table 5.2. Special cases of the reactions in Eq. 5.25

SoOSOH b
SOH
SOH
SOH
SOH
SOH

Ll-

Mm+

SRzsR
a

Na+
Na+
Cu2+
Pb z+
-

H+

Cl-

H+
H+

OHOH-

FUL-

P0,43Cit3-

Siloxane ditrigonal cavity.


e

Inorganic surface hydroxyl group.


Fulvic acid anion.

Citrate (2-hydroxypropane-l ,2,3-tricarboxylate).

zm-

i<

Rr -

pZp

"- 'i;"ftf5ttb

,-

?~~~~~;o..-

Reaction
SoO-(s) + Na+(aq) = SoONa(s)
SOH(s) + Na+(aq) = SONa(s) + H+(aq)
SOH(s) + Cqaq) + H+(aq) = SOHzCI(s)
SOH(s) + Cu2+(aq) + FUL"(aq) = SOCuFUL(s) + H+(aq)
SOH(s) + Pbz+(aq) + OH-(aq) = SOPbOH(s) + H''{aq)
SOH(s) + POl-(aq) + 2H+(aq) = SHZP04(s) + OlF'(aq)
SOH(s) + Cit3-(aq) + H+(aq) = SHCiC(s) = OH-(aq)

CHEMICAL MODELS OF SURFACE COMPLEXATION

167

ZSR - Zp.) Equation 5.25b describes the inner-sphere complexation of


an anionic species through the ligand exchange mechanism. In this
reaction, actually a special case of Eq. 5.25a, SR comprises a permanent
structural moiety S and an exchangeable ligand R; the entity R' in the
general surface complex SR'C does not exist, and so the symbol for the
complex is shortened to SC, where C == HxL q More generally, one could
substitute any cationic species (or set of species) for the proton in
Eq. 5.25b. Equations 5.25 can be modified without difficulty to describe
the case wherein several surface functional groups are involved simultaneously in a complexation reaction.V
The conditional equilibrium constant for the reaction in Eq. 5.25a can
be expressed in the form 13
(5.27)
where x is a mole fraction, ( ) is an activity in the aqueous solution phase,
and

An expression for "K can be derived in the context of the van der Waals
model with the help of Eq. 5.24 and the definition 10
p~vw)
Ni
xi =
'
=
- ~vw) - M
(5.28)
of the mole fraction of species i in a binary mixture. Equation 5.28 states
that the expected value of the mole fraction is equal to the ratio of the
relative probability that i exists in the mixture to the sum of relative
probabilities for all components in the mixture. With this equation and the
two explicit expressions for P}YW), one derives from Eq. 5.27 the equation
C

K -

exp[(JLc - ZSR'C e t/J~V;:~/ kBT](pzp)


QSR exp[(/LR - ZSR e t/J~i")/kBT](C)

QSR'C

(5.29)

According to thermodynamics, at equilibrium, 10

/Lc - /LR = /L[C(aq)] - /L[pzp(aq)]


=

/LO[C(aq)] - /LO[pZp(aq)] + kBT In

[(~~)]

(5.30)

where
/LO[C(aq)] == p/LO[Mm+(aq)] + q/LO[LI-(aq)] + Y/LO[OH-(aq)]
and /Lo is a Standard-State chemical potential. (In Eq. 5.30, when P exists,
/LR is taken to be the chemical potential per molecule of P in the solid
adsorbent.) The combination of Eqs. 5.29 and 5.30 produces the model
result
"K .. (K~R'C/ KSR)exp[ -(ZSR'C" e t/J~R~~' - ZSR e t/J~V;/ k n T] (5.31)

THE SURFACE CHEMISTRY OF SOilS

168

where

KSR'c = QSR'C exp(p,o[C(aq)]/ kBT)


K SR = QSR exp(JLO[pZp(aq)]/ kBT)

(5.32)

Finally, wi th the help of Eqs. 5.19a, 5.22, and 5.28, Eq. 5.31 Can be
written in the more explicitly composition-dependent form
cj(

= (KSR'c/ KSR)exp( + ZSR e l/JgR/ kBT)


x exp] - (ZSR'Cl/JgR'C + ZSR l/JgR) e XSR'C/ k B T]

(5.33)

According to the van der Waals model, the conditional equilibrium


constant fc1r the reaction in Eq. 5.25a is an exponential function of the
mole fracti.on of the surface complex species SR'C. A similar derivation
can be given for the conditional equilibrium constant pertaining to the
ligand exchange reaction in Eq. 5.25b by deleting R' and settingp = y =
and pZp = R r - in Eqs. 5.27, 5.30, 5.31, and 5.32.
The thefIllodynamic properties of the reaction in Eq. 5.33 follow
immediately from Eq. 5.27 and standard methods of quadrature.P For
example, atter equating the right sides of Eqs. 5.27 and 5.33 and forming
the common logarithm of all terms, one has the result

log (p Zp ) + Iog [XSR'C]


- - - Iog (C) -_ Iog [KSR'C. exp(+ZSR e l/JgR/kBT)]
XSR
K SR
(ZSR'C l/JgR'C + ZSR l/JgR)e
(In lO)k BT
XSR'C
(5.34)
Equation S.34 shows how measurements of the composition of the
adsorbent surface and of the activities in the aqueous solution phase can be
used to determine certain combinations of the constant parameters in the
van der w~als model. A graph of the left side of Eq. 5.34 versus XSR'C
should produce a straight line (at constant temperature) whose y intercept
and slope are related to the K and l/Jo parameters of Eqs. 5.22 and 5.32.
The standard quadrature formulav'
In K =

fa! In - dXSR'C

for the thertllodynamic equilibrium constant pertaining to Eq. 5.25a shows


that, in the van der Waals model,

(KSR'c/ KSR)exp[ -(ZSR'Cl/JgR'C - ZSRl/JgR)e/2k BT]

(5.35)

Therefore, the algebraic sum of the y intercept and one half the slope in
Eq. 5.34 is equal to the common logarithm of the equilibrium constant, K.
The rational activity coefficients of the two solid-phase species, SR and
SR'C, can also be calculated with standard quadrature formulas, J:l but it is

169

CHEMICAL MODELS OF SURFACE COMPLEXATION

sufficient here only to note that their ratio is


fSR'c
fSR

= cK = exp[-(ZSR'Co/SR'C - ZSRo/SR)ej2k BT]


x exp[+(ZSR'Co/~~~ - ZSRo/~V;)e/kBT]

(5.36)

according to Eqs. 5.31 and 5.35. Equation 5.36 shows that the exponential
factors in the van der Waals model equations for CK and K represent the
long-range coulomb contribution to the rational activity coefficients of the
surface species. The essence of the van der Waals model is that it provides
an estimate of how coulomb interactions among the surface species affect
the equilibrium constants for surface complexation reactions and the
activity coefficients of these species.
The van der Waals model can be generalized to include the situation in
which several kinds of surface complex coexist simultaneously or that in
which surface complexation involves polydentate ligancies with respect to
the surface functional groups (e.g., bidentate complexes with two groups
bonded to a bivalent metal cation). 12 These kinds of generalizations of the
model are not required explicitly in the present chapter, but their existence
underscores the broad utility of the van der Waals model as a conceptual
tool for elucidating the foundational aspects of surface complexation
theories.
5.3. THE CONSTANT CAPACITANCE MODEL

The constant capacitance mode114 ,15 is a molecular description of surface


complexation reactions involving the inorganic hydroxyl group. The chemical basis of the model can be developed from its three principal assumptions concerning the interfacial region:
1. Inorganic hydroxyl groups form only inner-sphere surface complexes
with adsorbed species.
2. The chemical reactions that describe surface complexation are

aSOH(s) + pMm+(aq) + qL1-(aq) + xH+(aq)


+ yOH-(aq) = (SO)aMp(OH)yHxLg(s) + aH+(aq)

(5.37a)

bSOH(s) + qL1-(aq) + xH+(aq) = SbHxLJ(s) + bOH-(aq)

(5.37b)

where 5 = pm + x - a - ql - y and 'Y = x + b - ql are valences of


the solid phase products. Equations 5.37 are the same as Eq. 5.25 (with
R = OH, ZSR = 0; P = H, Zp = +1; R' = 0, ZSR'C:; 5; Zsc:; 'Y,
r = 1) when b, the number of moles of reactive OH groups, equals
unity. In the constant capacitance model, the conditional equilibrium
constants for the reactions in Eq. 5.37 use the Constant Ionic Medium
Reference State for the activity coefficients of the aqueous species. 16
Because of this convention, the species Mm+(aq) and U-(aq) in
Eq. 5.37 never refer to the metal cation and the ligand making up the

170

THE SURFACE CHEMISTRY OF SOILS

background electrolyte. By analogy with Eq. 5.27, the conditional


equilibrium constant for the reaction in Eq. 5.37a can be written in the
form
x
[H+:r
C K I = _s,-,o;..:c=--__
(5.38)
xSOH [C]
where
and the square brackets refer to concentration in moles per cubic
decimeter. Equation 5.38 is written under the assumption that the
concentrations [H+] and [C] are enough smaller than the concentration
of the background electrolyte to justify setting the aqueous phase
activity coefficients equal to 1 on the scale of the Constant Ionic
Medium Reference State.l"
3. When the total particle charge is small in absolute magnitude, it is
proportional to the inner potential at the particle surface:
<Tp

c,

(5.39)

where <Tp is the surface density of total particle charge (Eq. 3.2), C (in
faradays per square meter) is a differential capacitance density, and t/Js is
the inner potential at the surface to which <Tp refers. Equation 5.39 is
consistent with the balance of surface charge, Eq. 3.3a, if <TD is
calculated with the DDL theory expression in Eq. 5.15. Upon expanding the right side of Eq. 5.15 to second order in a MacLaurin series in
the variable t/J, one finds the approximate relationship

~CjZ7T/2t/J

<TD = - [(E oDF2/ R T)

(5.40)

which is consistent with Eqs. 3.3a and 5.39 when t/J == t/Js and
IZjFt/Js!RTI is small enough to justify replacing the exponential function in Eq. 5.15 by its first three terms in a MacLaurin expansion.
In the constant capacitance model, the surface
complexation reactions that involve H+ or OH- alone are special cases of
Eq. 5.37a: 14
SOHt(s) = SOH(s) + H+(aq)
(5.41a)
THE NET PROTON CHARGE.

SOH(s) = SO-(s) + H''{aq)

(5.41b)

The conditional equilibrium constants for these two reactions are 15


KS

al -

xsoH[H+]
+

XSOH 2

K" _ "so-[H+]
liZ

(5.42a)

" SOH

(5.42b)

CHEMICAL MODELS OF SURFACE COMPLEXATION

171

(Note that Eq. 5.41a is written in the reverse of the special case of
Eq. 5.37a corresponding to a = x = 1, P =q = Y = 0, and that K~l in
Eq. 5.42a is the inverse of "K' in Eq. 5.38 applied to this special case.)
Besides the conditional constants in Eq. 5.42, the constant capacitance
model specifies two intrinsic equilibrium constants for the proton
reactions'f
K~l(int)

= K~l exp( - Ft/lsiRT)

(5.43a)

K~2(int)

(5.43b)

K~2

exp( -Ft/lslRT)

The intrinsic equilibrium constants are postulated to be independent of the


composition of the solid phase (although they remain conditional in the
sense of the Constant Ionic Medium Reference State). They can be
determined experimentally on the basis of the linear expressions that result
from combining Eqs. 5.39, 5.42, and 5.43:
XSOH )
[+] _
s .
F
-log ( xSOHi - log H - -log Ka1(mt) - (In 10)CRT
-log[H+] - 10g(:::J

-log

K~2(int) -

(In

1~CRT

UH

UH

(5.44a)
(5.44b)

where up has been set equal to UH, the net proton surface charge density,
because of Eq. 3.2. Equations 5.44 can be applied to proton titration data
under the assumption that'"
pH < PZNPC
pH> PZNPC

(5.45)

where U max == FM/N A is the maximum absolute value of UH, M is the


total number of reactive OH groups per unit area of adsorbent,
N A = 6.023 X 1023 mol"! is the Avogadro constant, and PZNPC is the
point of zero net proton charge (Sec. 3.2). Equations 5.45 are equivalent
to assuming that the species SO-(s) does not exist below the PZNPC and
the species SOHt(s) does not exist above the PZNPC; i.e., in general,
(5.46)
according to Eq. 1.24. The combination of Eqs. 5.44 and 5.45 permits the
calculation of -log K~i(i = 1 or 2) and the parameter C from the y
intercept and slope of a plot of the left side of Eq. 5.44 against either xSOHi
or XSo-, as illustrated in Fig. 5.2 for y-Alz0 3 Y Once the values of the
common logarithms of K~l(int) and K~2(int) and of the capacitance density
C have been determined in this way, they can be used with Eqs. 5.39,5.42,
and 5.43 to calculate UH and the distribution of SOH, SOHt, and SO- at
any pH value.
The conformity of the proton titration data for ,...A1 203 to the constant
capacitance model (within experimental precision) is evident in Fig. 5.2.

0'

-log K~I (int) = 7.2

_o~

~o. ~

y-A1 20 3

.~

If

~
-I

O.IM NaCI04

10.0

6-

9.5~
:::: log

21
0

0.08

0.24

0.40

0.56

XSOH+2

0.72

9.0
0

--- .
6-

0.08

-Kg

6-

6-

6-

6- 6- 6- 6-

6-

(int) =9.5

0.24

0.40

0.56

XSO-

Ipre 5.2. Plots of the left side of Eq. 5.44 against either xsoH! or XSo- for y-A1203 suspended in 0.1 M NaCI04 (3.2 kg'm- 3
IISpCDSion).17 The different symbols represent separate experiments.

__~

'~

- ".~ -~
;;;...,~.- ~_="-.."":;,,

-.

- - .-

'.

<:..::0- ....... - --

6-

173

CHEMICAL MODELS OF SURFACE COMPLEXATION

Equations 5.44 can be recognized as special cases of the van der Waals
model expression in Eq. 5.34 obtained by setting ZSR 0 and IZsR'c1 == 1.
Equation 5.44a corresponds exactly to Eq. 5.34 after the identifications

K:al (iInt)

=
-

SO H
KK

SOH!

=
-

CTmax
0
l/JSOH!

(5.47a)

have been made, and the same kind of result follows for Eq. 5.44b with the
definitions
S'

Kdmt)

K so K SO H

=--

U max

C=--o-

l/Jso-

(5.47b)

where tIP is the van der Waals potential. These correspondences show that
the intrinsic equilibrium constants are related closely to the partition
functions of the surface species (Eq. 5.32) and that the capacitance factor
is related to the ratio of the maximum absolute value of CTH to the
maximum absolute value of the van der Waals mean electric potential
(Eqs. 5.22 and 5.45). Thus the intrinsic equilibrium constants provide a
quantitative measure of the strength of the chemical bonds between a
surface OH group or a surface complex and the remainder of the solid
adsorbent. The capacitance parameter is seen to be the capacitance per
unit of adsorbent surface area associated with either a fully protonated or a
fully dissociated hydroxyl surface. This relationship is consistent with the
special case of Eq. 5.39 that occurs when ICTHI = CTmax and ll/Jsl is equal to its
maximum value.
The intrinsic equilibrium constants have thermodynamic significance, in
that they determine the PZNPC through the condition XSo- = XSOH!
(pH = PZNPC) applied to Eq. 5.42: 14
PZNPC =

(pK~1(int)

+ pK~2(int

(5.48)

However, the intrinsic constants are not the same as the thermodynamic
equilibrium constants for the reactions in Eq. 5.41. The thermodynamic
constants can be calculated with the help of Eq. 5.35 and the correspondences in Eq. 5.47:

K1 =

K~1(int)exp(FCTmax/2CRT)

K2 == K: 2(int)exp( -FCTmax/2CRT)

(5.49a)
(5.49b)

where Ki(i = 1,2) is a thermodynamic constant. Equations 5.49 reflect the


fact that the intrinsic equilibrium constants refer to a Standard State in
which the species SOH{(s) or SO-(s) are at unit mole fraction but create
no surface charge density,12.14.18 i.e., K1 = Kai(int) (i = 1,2) only if
O'max = O. This unconventional Standard State deprives the intrinsic constants of ordinary thermodynamic significance, although they are directly
proportional to thermodynamic cquilibrlum constants.

174

THE SURFACE CHEMISTRY OF SOilS

Figure 5.2, with its two lines of unequal slope (in absolute value),
illustrates another typical feature of applications of the constant capacitance model. The value of the capacitance parameter C inferred from each
slope is not the same above and below the PZNPC. 14,15,17 This result is in
conflict with the nonspecific, coulombic nature of t/J~vw) indicated in
Sec. 5.2; with the DDL theory relation between aD and t/Jin Eq. 5.40; and
with the general thermodynamic requirement that t/JgOH! = It/Jgo-I in a
ternary system comprising SOH(s), SOHt(s), and SO-(S).19 Therefore,
the lack of uniformity in the experimental value of C for pH values above
and below the PZNPC must be regarded as an inherent shortcoming of the
constant capacitance model.
METAL CATION ADSORPTION. The formation of inner-sphere surface complexes involving metal cations is typically described in the constant capacitance model with the chemical reactions 15,17,20
SOH(s) + Mm+(aq) = SOM(m-1)(s) + H+(aq)
2S0H(s) + Mm+(aq)

(5.50a)

= (SOhM(m-2\s) + 2H+(aq) (5.50b)

The conditional equilibrium constants for these reactions are 15

{SOM(m-1)}[H+]
*Ki == {SOH}[M m+]

s _

f32

{(SOh M(m-2)}[H+]2
{SOH}2[M m+]

(5.51a)
(5.51b)

where { } denotes concentration in moles per kilogram of adsorbent.


Equations 5.50a and 5.51a are special cases of Eqs. 5.25a and 5.27,
respectively, obtained by setting pZp == H+, [C] == [M"'"], SR = SOH,
and SR'C == SOM. The conditional constants *Ki and *f3'2 are usually
evaluated by graphic methods that employ measured concentrations of M
in surface complexes and calculated values of {SOH} based on experimental determinations of the acidity constants K~l and K~ .15,17,20
If M is a metal in oxidation state II, then both surface species in
Eq. 5.50b are electrically neutral and the van der Waals model predicts
that *f3'2 is independent of surface charge, i.e., that *f3'2 = *f32(int).
However, since the surface complex SOM+ is positively charged, the
definition'f
*KWnt)

= *Ki exp(Ft/JslRT)

(5.52)

should be applied by analogy with Eq. 5.43. In practice,15,17,20 it is found


that *Ki either does not depend significantly on the total particle charge or
shows a much smaller dependence than the exponential one in Eq. 5.52.
Unless this result is caused by a lack of sensitivity in the methods used to
determine *Ki, it implies an additional problem of self-consistency in the
constant capacitance model.

175

CHEMICAL MODELS OF SURFACE COMPLEXATION

100

o
w

Cu(II)

CO

0:::

g
o

60

<{

Z
W

()

0:::

SILICA GEL
3M NaCI0 4

Q.

pH
Figure 5.3. Adsorption edges for Fe (III) , Cu(II), Cd(II) on silica gel suspended in
3 M NaCI0 4 . The solid lines represent calculated values based on the constant
capacitance model. (After Schindler et al. 20 )

Despite these difficulties with internal consistency, the constant capacitance model can provide an excellent quantitative description of the dependence of metal adsorption on pH, as exemplified in Fig. 5.3., and it gives
a prediction of the pH dependence of 0" p in the presence of adsorptive
bivalent meta! cations that is in complete qualitative agreement with the
surface charge behavior implied in Fig. 4.7. 15 The increase in metal
adsorption with increasing pH value occurs in the constant capacitance
model solely because of an increase in the concentration of the 80- species
and in the strength of the inner-sphere complexes between this species and
the metal cation, as reflected in the conditional constants in Eq. 5.51.
Hydrolysis of the adsorbed metal cation is not usually invoked in the
constant capacitance model because significant adsorption occurs at pH
values well below the PZNPC of the adsorbent and below the onset of
hydrolysis of the adsorptive metal in the aqueous solution phase. 14,15 With
respect to the reversals of surface charge implied in Fig. 4.7, the constant
capacitance model provides a mechanism based on a competition between
the reactions in Eq. 5.41 and those in Eq. 5.50, In the absence of an
adsorptive metal cation, the model predicts a single charge reversal at the
PZNPC (cf. the lowest curve in Fig. 4.7). When the metal cation is
present, an increase in pH value contributes to a negative surface charge

THE SURFACE CHEMISTRY OF SOilS

176

through Eq. 5.41b and a positive surface charge (for m > 1) through
Eq. 5.50a. If the concentration of metal cations in the aqueous solution
phase and the value of *K~ are large enough, the reaction in Eq. 5.50a
eventually dominates and the surface charge increases with increasing pH
value, as indicated in Fig. 4.7. However, as the pH value becomes high
enough to induce significant hydrolysis of the adsorptive metal cation in the
aqueous solution phase, the surface charge again decreases, according to
the constant capacitance model, because the concentration of the
SOM(m-1) species is reduced sharply by the decline in the value of [M"."]
in Eq. 5.51a. This mechanism of surface charge decrease does not
predict that the total surface charge at high pH values will emulate that of a
hydroxy-polymer coating of the adsorbed metal cation; instead it predicts
that the change at high pH values will be similar to that expected in the
absence of the adsorbed metal. 21
The constant capacitance model postulates that
anions react with surface hydroxyl groups through the ligand exchange
mechanism embodied in Eq. 5.37b. 14 ,22 In typical applications, the value
of b in Eq. 5.37b is either 1 or 2 and the corresponding conditional
equilibrium constants have the form
ANION ADSORPTION.

(5.53a)

s =

{S2HxL~Y+1)}[OH-Y

f32 - {SOHf[HxL~X-ql)]

(5.53b)

where in each case the equilibrium constant for the formation of


HxL~-ql)(aq) from H+(aq) and L'-(aq) has been incorporated. By
analogy with Eq. 5.43, an intrinsic equilibrium constant that corresponds
to K~ can be defined as
Ki(int) ==

K~

exp( "IFl/JsI RT)

(5.54)

and the analog of Eq. 5.44 can be derived:


10g[OH-] + 10g(XSHL) .
XSOH
=

log Ki(int) - (In

10g[HxL~-ql)]

1~;CRT up

(5.55)

If experimental conditions are arranged such that SOH and SHxL q are the
only surface species, then up = "I U max XSHL, where XSHL is the mole

fraction of SHxL q, and Eq. 5.55 becomes a special case of the van der
Waals model expression in Eq. 5.34. In particular, the identifications
pZp _ OH-, SR'C - SHxL q , SR - SOH, ZSR = 0, ZSR'C = "I. and

177

CHEMICAL MODELS OF SURFACE COMPLEXATION

[C] == [HxL q(x-ql)] can be made. It also follows that


KWnt) == K SH L

C ==

K SO H

'Y~max
~SHL

(5.56)

in complete analogy with Eq. 5.47.


The constant capacitance model has been applied successfully to describe the adsorption of both inorganic and organic anions by hydrous
oxides.F Both adsorption isotherms and adsorption envelopes like those
illustrated in Fig. 4.8 are accounted for quantitatively, except in the case of
sulfate adsorption. It is possible, however, that this anion adsorbs through
outer-sphere instead of inner-sphere complex formation, as discussed in
Sec. 4.4. Aside from sulfate, anion adsorption envelopes are predicted in
the constant capacitance model on the basis of an increase in [OH-],
which causes the concentrations of the surface complexes in Eq. 5.53 to
decrease, and a decrease in up'

5.4. THE TRIPLE LAYER MODEL

The triple layer modee3 ,24 offers a molecular description of surface


complexation reactions that differs from the constant capacitance model in
several fundamental respects. These differences can be brought into clear
relief through a comparative listing of the principal chemical assumptions
that underlie the triple layer model:
1. The proton and the hydroxide ion form inner-sphere surface complexes.

All other adsorbed metal cations and inorganic or organic anions form
outer-sphere complexes.
2. If the reacting surface functional group is an inorganic hydroxyl group,
the chemical reactions that describe surface complexation are special
cases of Eq. 5.37a. However, except for the protonation-proton
dissociation reactions in Eq. 5.41, the solid phase product,
(SO)aMp(OH)yHxL;(s), is understood to always represent an outersphere complex. Therefore, when metal cations or cationic complexes
adsorb, a water molecule separates them from the (SO)a unit. When
anionic complexes adsorb, they, too, are separated from the protonated
(SO)a unit by a water molecule. The ligand exchange mechanism of
Eq. 5.37b is not invoked as the basis for surface complexes with anions.
Conditional equilibrium constants in the triple layer model use the
Infinite Dilution Reference State for the activity coefficients of aqueous
species. 16 With this convention, the species M m + (aq) and L l - (aq) in the
reaction in Eq. 5.37a can refer to the metal cation and the anion of the
background electrolyte. The conditional equilibrium constant for this
general reaction takes the form
cK

= "sodH + )a
"~Cl"(C)

(5.57)

THE SURFACE CHEMISTRY OF SOilS

178

which is the special case of Eq. 5.27 in which SR' = SO, pZp = H+,
and SR = SOH.
3. The relationship between surface charge and inner potential is specified
through the following equations:

O'd

O'H = C1(o/s - 0//3)

(5.58a)

O'd = CZ(o/d - 0//3)

(5.58b)

-sgn(o/d){ 2eoDRT:t Cj[exp(-ZjFo/d/RT)- 1Jf/Z (5.58c)

where C 1 and Cz are integral capacitance densities and the inner


potentials o/s, .0//3, and o/d are identified with the three planes labeled
"s," "{3," and "d" in Fig. 5.4. Equation 5.58c is the same as Eq. 5.15
applied to the d plane, i.e., DDL theory is assumed to apply to the
aqueous species that neutralize surface charge, with the diffuse region
beginning at the outer periphery of the water molecules that solvate
ions adsorbed in outer-sphere surface complexes. It is evident from
Fig. 5.4 that the net charge density in the {3 plane, 0'/3' corresponds to

Figure 5.4. A schematic portrayal of an inorganic hydroxyl surface, showing


planes associated with surface hydroxyl groups ("s"), inner-sphere complexes
("a"), outer-sphere complexes ("{3"), and the diffuse ion swarm ("d").

s a

OXYGEN

CHEMICAL MODELS OF SURFACE COMPLEXATION

179

and that surface charge balance (Eq. 3.36) is expressed mathematically by the equation
(TOS

(5.59)
in the triple layer model. By contrast with the constant capacitance
model, the linear relationships in Eqs. 5.58a and 5.58b are assumed
valid whatever the magnitudes of the inner potentials, 0/.. o/f3, and o/d'
Moreover, the charge-balance condition, Eq. 5.59, is imposed explicitly
and the full DDL theory expression for (T d ==' (TD is used. Note that
Eqs. 5.58 and 5.59 are not consistent for arbitrary values of the surface
charge densities and inner potentials unless ions are present in the {3
plane under all circumstances.
THE NET PROTON CHARGE. In the triple layer model, the protonation and
proton dissociation reactions in Eq. 5.41 are described by the conditional
equilibrium constants
xSOH(H+)

Qal == -::'=,'-'----'--

(5.60a)

XSOHt

Qa2 ==

xso-(H+)

(5.6Ob)

--=.=---..;'--...:..

XSOH

Besides the use of the Infinite Dilution Reference State with respect
to the activity of H+(aq), the Qa,U = 1,2) in Eq. 5.60 differ from the
K~, U = 1,2) in Eq. 5.42 through the stipulation that XSOHt and xsoare mole fractions of the uncomplexed solid phase species SOHt(s) and
SO-(s). Thus complexed species like SOH 2L<1- 1)(S) and SOM(m-l)(s) are
excluded from Eq. 5.60 even though SOHt and SO- form part of their
structure on the molecular level. Instead, special cases of Eq. 5.37a, such
as Eq. 5.50 and the sum of Eqs. 4.44a and 4.44c, are assumed to apply to
the background electrolyte and contribute to (TH' The conditional equilibrium constants for these latter reactions are, for the common example of a
1: 1 background electrolyte ML(aq)

*QL- == xsoH(H+)(L-)

(5.6Oc)

XSOH2L

*Qw== XSOM (H+)


XSOH (M+)

(5.6Od)

Corresponding to these four conditional constants are the triple layer


model intrinsic equilibrium constants."

K::r

iIE

Qal exp(-eo/s/kBT)

"'K~l. "'Qa exp[-e(lP" - o/(3)/kBT]

(i

1,2)

(5.61a)

(a = L - ,M+) (5.61b)

The four intrinsic constants are postulated to be independent of the

180

THE SURFACE CHEMISTRY OF SOilS

composition of-both the solid and the aqueous solution phase. They can be
determined experimentally on the basis of linear expressions that follow
from the combination of Eqs. 5.58a, 5.60, and 5.61:
,
(XSOH)'
,
e"'f3
pH -log'
,
= -logK~~t
- e
UH XSOHt
(In 1O)C1kBT
(In 1O)kBT

pH-log

(XSO-) _ -logKint
XSOH

a2 -

e
(In1O)C

pH -IOg(x::::J -log(L-) = -log *


pH _IOg(x

SO M

1kB

TuH

e"'f3
(Inl0)k T
B

Kt~ - (In 1O)eC~::'~ u~'>:

(562a)
.

(5.62b)
(5.62c)

+ 10g(M+) = -log * Kk1\ -

XSOH

where

UH = (FM/NA)(xSOHt +

XSOH 2L -

xso- -

XSOM)

(5.63)

Under the assumption that experimental conditions have been arranged to


make one of the surface species that contribute to UH dominant, precise
graphic extrapolation methods can be applied to determine the intrinsic
equilibrium constants in Eq. 5.62 from measurements of the conditional
constants and UH over a range of concentrations of the background
electrolyte. 25
The linear relationship implied in Eqs. 5.62c and 5.62d is illustrated in
Fig. 5.5 for rutile suspended in LiN03 . 26 It is assumed in this application
that the concentration of LiN0 3 and the pH are high enough to justify the
approximation of neglecting all mole fractions except XSOM in Eq. 5.63.
The y intercept of the line through the data equals 7.2 and corresponds to
-log *Kb\, according to Eq. 5.62d. The slope of the line equals 25 and
therefore

_.
eFM
_ 1
.-2
Ct - 4'25(ln 1O)NAkB T'-- .30 F m
1""_,,,,1

at 298.15 K, where M = 1.2 X 1019 m- 2 has been used in the calculation.r'' These two parameters can be related to the van der Waals
models parameters by setting SR'C = SOM and SR = SOH in Eq. 5.34
and comparing the resulting expression with Eq. 5.62d. In this application,
ZSOH = 0 but ZSOM = 1 because the van der Waals average potential is
created by the solvated M+ cations in the f3 plane. Therefore, with the
correspondences pZp = H+ and (C) = (M") inserted into Eq. 5.34, the
triple layer model parameters can be expressed by the equations'f
*Kint _
M+ -

KSO M
K SO H

1 -

FM
0

N A "'SOM

(5.64)

where "'~OM is the maximum van der Waals electric potential created by

CHEMICAL MODELS OF SURFACE COMPLEXATION

181

I I r-----,.----r---r-----,-----,-----,-----,----.-..,

Rutile in LiN03

+
...J

*o0LiT
.. 1M

O.IM

__

5 ~

__JL.___

0.02

__JL.___

___I_ ____I_ ____I_ _.......L_ _.......L_..J

0.04..

0.06

0.08

-uH I(FM/.A)

Figure 5.5. A graph of -log *Qu+ (Eq. 5.60d) versus -uH/(FMNA ) (Eq. 5.63)
for rutile suspended in LiN0 3 . The line through the data conforms to Eq. 5.62d.
(After Davis et al. 23)

the complexed M cations and FM/N A is the maximum absolute value of


(TH' The molecular interpretation of Eq. 5.64 is similar to that of Eq. 5.47,
with the parameter C 1 being the capacitance, per unit of adsorbent surface
area, of an array of solvated M+ complexed by SO-.
The thermodynamic equilibrium constant for the reaction
SOH(s) + M+(aq) = SOM(s) + H''{aq)

(5.65)

can be calculated in the triple layer model with the help of Eqs. 5.35 and
5.64:
(5.66)
Just as in the constant capacitance model, the intrinsic equilibrium
constant in the triple layer model refers to an unconventional standard
state wherein the complexed species create no coulombic effects, i.e.,
*K~~ = K when c/1~OM = O.
Equations 5.62c and 5.62d provide the basis for two independent
determinations of the .capacitance parameter C 1 in the triple layer model.
However, the check on internal consistency that these two separate
evaluations of C, could furnish has never been performed in any published
study.2:l,24,27-.,o In fact, C 1 has been taken universally as an adjustable

182

THE SURFACE CHEMISTRY OF SOilS

model parameter along with the truly adjustable capacitance factor Cz in


Eq. 5.58b. For this reason, no conclusion is possible as to the accuracy of
the triple layer model in self-consistently representing the molecular
structure of the inorganic hydroxyl surface, notwithstanding its widespread
success in reproducing surface charge data for hydrous oxides suspended in
electrolyte solutions. Z4
The formation of outer-sphere surface complexes involving metal cations has been described typically in the triple
layer model by the reactions z3,z4
METAL CATION ADSORPTION.

SOH(s) + Mm+(aq)
SOH(s) + Mm+(aq) + H zO(l)

SOM(m-l)(s) + H+(aq)

(5.67a)

= SOMOH(s) -+ 2H+(aq) (5.67b)

Equation 5.67b is obtained from Eq. 5.25a (SR = SOH, ZSR = 0,


P = Y = 1, q = x = 0, SR' = SO, ZSR'C = 0, pZp = H+) by adding the
ionization reaction of water. The conditional equilibrium constants for
these reactions are Z3
(5.68a)
(5.68b)
Reactions involving two SOH to form (SOhM(m-Z) are excluded because
. they invariably lead to serious disagreement between measured and
calculated adsorption edges for metal cations. z3,z4 Moreover, the values of
* KMm+ determined in applications of the triple layer model usually are
small enough to preclude the dominance of the outer-sphere complex
SOM(m-l) except at pH values well below the PZNPC. This behavior is in
direct contradiction with that typically observed in applications of the
constant capacitance model, where the inner-sphere complex SOM(m-l)
tends to be dominant even above the PZNPC. The conditional constants in
Eq. 5.68 are related to triple layer model intrinsic equilibrium constants
through the equationz3,z4
*K~nt

==

*Ka exp[ -e(l/Is - (m - l)l/I~)/kBT]

(a = Mm+,MOH(m-l

(5.69)

When M'?" represents a metal cation not in the background electrolyte,


the intrinsic constants are determined by fitting the triple layer model to
adsorption edge data. This fitting entails a surface speciation calculation
with previously measured values of the intrinsic constants in Eq. 5.61, the
capacitance parameters C\ and C2 , and the parameter M. The computation includes Eqs. 5.58, 5.59, and 5.69, as well as surface charge and mole
balance equations imposed as constraints. 23 31

183

CHEMICAL MODELS OF SURFACE COMPLEXATION

The triple layer model has been shown to provide quantitative descriptions of both adsorption edges and uwpH relationships for a variety of
hydrous oxide adsorbents suspended in aqueous solutions ranging in
composition from single, 1: 1 electrolytes to major-ion seawater. 23,24,27-30
The increase in adsorption with pH commonly observed for bivalent metal
cations (Fig. 4.6) is interpreted in the triple layer model as the result of an
increase in the concentration of the SO- species and in the formation of
the outer-sphere complex SOMOH. This effect is illustrated in Fig. 5.6 for
the adsorption edge of Pb(II) on y-A12 0 3 suspended in 0.1 M NaCI0 4 . 32
The model curve, which follows the experimental data closely reflects
almost entirely the contribution of complexed PbOH+ cations, since the
SOPb + species never accounts for more than 20 per cent of the adsorbed
lead and the PZNPC of y-A120 3 is about 8.3. Thus a kind of surfaceinduced hydrolysis of the adsorptive metal cation plays a major role in the
triple layer model, whereas in the constant capacitance model it plays no
role at all. This difference is made all the more striking by the fact that the
data in Fig. 5.6 have been described with equal quantitative accuracy by
the constant capacitance model. 17 ,33
The charge reversal behavior shown in Fig. 4.7 also can be described by
the triple layer model. 34 The mechanism relies on the competition between
the reactions in Eq. 5.41 and that in Eq. 5.67b. When an adsorptive,
bivalent metal cation is present, an increase in pH value causes the

Figure 5.6. The adsorption edge for Pb(II) on y-Al z0 3 suspended in 0.1 M
. NaCl0 4 The solid line represents the contribution of SOPb+ alone. (After Davis
and Leckie z3)
100 r------r-----r----.----r---r---a::::;==-----,

80
"0
CI.l
X
CI.l

a.
E

.""

60

50- _Pb 2+ and 50- - PbOH+

-...
cCI.l::

40

CI.l

0..

20
_____ - - - - -

--------,-50--Pb 2 + only

......- -....
7.0

OL.--.....I--.....L--.....I.--.....I..--~--

3.5

4.0

4.5

5.5

5.0

pH

6.0

6.5

THE SURFACE CHEMISTRY OF SOilS

184

formation of the SOMOH species to increase sharply. At a given pH value,


this species increases the particle charge (i.e., a decrease of the net charge
in the d plane) relative to its value when the adsorptive, bivalent metal
cation is absent. This point is made clearly through a consideration of
charge balance in Eq. 5.59: at a fixed pH value UH remains constant while
uf3 increases when MOH+ appears in the (3 plane. The increase in particle
charge through outer-sphere surface complexation can be large enough, if
* Kkitow is large enough, to reverse the negative particle charge expected
from an increase in uncomplexed SO- as the pH value increases. At very
high pH values, if the concentration of MOH+(aq) is reduced through the
formation of more hydrolyzed species, e.g., M(OH)g(aq), there could be a
loss of SOMOH on the adsorbent and a return of the particle charge to
negative values. Like the constant capacitance model, the triple layer
model does not predict that the ultimate particle charge will emulate that
of a hydroxy-polymer coating of the adsorbed, bivalent metal cation.
The triple layer model postulates that anions react
with surface hydroxyl groups according to the general equation
ANION ADSORPTION.

aSOH(s) + L1-(aq) + xH+(aq)

= (SO)aHxL 8(S) + aH+(aq) (5.70)

which is a version of Eq. 5.37a. In most applications.P the stoichiometric


coefficient a in Eq. 5.70 is set equal to unity but x can vary from 0 to I - 1.
The conditional and intrinsic equilibrium constants for the surface complexation reactions have the form
(5.71a)
(5.71b)
in analogy with Eqs. 5.6Oc and 5.61b. The values of *K~:L are determined
by fitting the model to adsorption envelope data following procedures
similar to those outlined for metal cation adsorption.P A striking confirmation of Eq. 5.70 for SOH on goethite and L = CI or CI0 4 has been
found through a kinetic study of the adsorption of these two anions in
acidic suspensions" The kinetics data revealed that, after protonation of
the SOH group, the anion diffuses to the {3 plane, where it is bound
temporarily as a species with two-dimensional mobility. Shortly thereafter,
the anion moves about in the {3 plane until it locates an SOHt site where it
can form an outer-sphere complex. The initial protonation step involves a
time constant of the order of milliseconds, the diffusion of the anion to the
{3 plane involves a time constant near 0.1 JLS, and the surface complexation
reaction involves a time constant in microseconds. The value of '" KWb
calculated from the rate constants" for the surface complexation reaction
agreed with the static value determined from adsorption experiments. 24
Reasonable agreement with experimental adsorption envelopes, like those

CHEMICAL MODELS OF SURFACE COMPLEXATION

185

in Fig. 4.8, has been obtained even when the triple layer model is applied
to strongly adsorptive oxyanions. However, except for SO~-, this agreement is likely to be a model artifact since the mechanism of strong
oxyanion adsorption is ligand exchange, as discussed in Sec. 4.4.
5.5. THE OBJECTIVE MODEL

The objective model'" derives its name from a pluralistic conceptualization of the species possible on an inorganic hydroxyl surface. This conceptualization, as will emerge clearly from the following list of model
hypotheses, includes features of both the constant capacitance and the
triple layer model.
1. Surface complexation involves all four of the planes depicted in

Fig. 5.4. Complexed protons and hydroxide ions reside in the s plane,
inner-sphere complexes containing trace metal cations or oxyanions are
assigned to the a plane, outer-sphere complexes with the ions of a
background electrolyte are assigned to the 13 plane, and the d plane
marks the beginning of the aqueous solution phase, where the diffuse
ion swarm is found.
2. The activity of a species i located in the plane A(A = s,a, or 13) is equal
to the ratio of the surface excess of i (Eq. 4.5) to moles of i per unit area
that can still be accommodated in the plane A:36
r~w)
I

(5.72)

where ail.. is the activity of i in the plane A, r}w) is the surface excess of i,
M A is the maximum number of surface species possible per unit area in
the plane A, and the sum is over all species adsorbed in the plane A. The
activity of species i in the plane Ais assumed further to be related to the
activity in the aqueous solution phase through the equatiorr"
(5.73)
where K'A is an empirical parameter representing the interaction
between i and a surface functional group, a, is the activity of i in the
aqueous solution phase, Z, is the valence of the adsorbed species, and
!/JA is the inner potential at the plane Aexpressed relative to a potential
in the aqueous solution phase far from the surface. The combination of
Eqs. 5.72 and 5.73 provides the basic adsorption equation of the
objective model:

r}W)

= [(MAIN..... ) -

:t rjW)]

KIA

a, exp(-Z,F !/JAIRT)

(5.74)

Equation 5.74 represents a set of coupled algebraic equations (one for


each adsorbed species in the plane A) that is subject to the mole-balance

186

THE SURFACE CHEMISTRY OF SOILS

constraint
(5.75)
3. The relationships between surface charge density and inner potential
are specified by expressions similar to Eq. 5.58:
UH

= G sa( l/Js

- l/Ja)

UH

+ Ua = Gaf3 (l/Ja - l/Jf3)

(5.76a)
(5.76b)

where the G parameters are capacitance densities and U a is identified


with UIS in Eq. 3.2. The surface charge density Ud is also given by
Eq. 5.58c as a function of l/Jd' and the balance of surface charge takes
the form of Eq. 3.3:
(5.77)

Equations 5.58c, 5.76, and 5.77 define the charge-potential relations in


the objective model for any set of adsorbed species in any plane or in
the diffuse ion swarm.
Although surface complexes are considered in the objective model in the
spirit of Fig. 5.4, no complexation reactions or equilibrium constants are
used. Instead, the activity relations in Eqs. 5.72 and 5.73 are invoked to
.describe complex formation. Equation 5.72 is identical with the equation
for the activity of an adsorbed species that experiences no lateral interactions on a surfacer" On the other hand, Eq. 5.73 can be derived by
applying Eq. 3.20 to an adsorbed species under the assumption that the
inner potential, cf>, can be equated to l/JA:

jL[i] = go + RT In aiA + ZjFl/JA

(5.78a)

ji[i] = JL[i(aq)] = JL[i(aq)] + RT In a,

(5.78b)

and (d. Eq. 5.30)

for equilibrium with respect to the transfer of species i across the interfacial
region. Equation 5.73 follows from Eq. 5.78 after making the definition

KiA == exp{(JLO[i(aq)] - go)/RT}


As pointed out in Sec. 3.3, Eq. 5.78a has no thermodynamic significance
when l/JA is an inner potential and Eq. 5.73 must be regarded as an ad hoc
model assumption.

Equation 5.72 can be applied successively to the species SOH and


SOHt in the s plane to derive the expression
r(w) a
K SO H S
K~1 = ;~~ w = K
exp(Fl/Js/RT)
(5.79)
SOH!

SOHiS

where K SO H S ... exp(-o/RT). Equation 5.79 is the same, formally, as


Eq, 5.43a and is a special case of the van der Waals model expression in

CHEMICAL MODELS OF SURFACE COMPLEXATION

187

Eq. 5.34. Thus the protonation (and proton dissociation) of the inorganic
hydroxyl surface is described in the objective model just as it is in the
constant capacitance model. However, in the objective model, aH is given
the explicit mathematical representation
- F(r(w)
r(w) ) aH SOH~ so- = as
_ (Ms / NA)[KsoH}saw exp( -F/Js/ RT) - Kso-saoH- exp(F/Js/ RT)]
1 + KSOH~SaWexp(-F/Js/RT) + KSO-SaOH-exp(F/JslRT)
(5.80)
which follows from Eqs. 5.74 and 5.75 applied to SOH, SOHt, and
SO-.36 Equation 5.80 is a special case of the surface charge density
relationship applied to each plane A:
aA = F ~

zjrt)

(A = s,a, or f3)

(5.81)

where the sum extends over all species j in the A plane. Equations 5.81 are
subject to the constraints implied in Eqs. 5.58c and 5.77.
The objective model considers the parameters MA , KiA' and GA>..'
(A,A' = s, a, or f3; A 1= A') to be adjustable curve-fitting parameters. The
constants pertaining to the sand f3 planes are determined by fitting the
model to potentiometric titration data (aH as a function of pH) obtained in
the presence of a background electrolyter" An additional check on this fit
'can be made if one assumes that /Jd is the same as the zeta potential and
measured values of ( as a function of pH are available. Once the
parameters for the sand f3 planes have been established, adsorption
isotherm, adsorption edge or adsorption envelope, and surface charge data
pertaining to the reaction of the inorganic surface hydroxyl group with
trace metal cations or with oxyanions can be used to determine the
parameters in the a plane from curve-fitting algorithms.39
It should be apparent even from this brief discussion that the objective
model is essentially a parameter-optimization procedure for speciating the
interfacial region near an array of inorganic surface hydroxyl groups. The
model exhibits some features of a hybrid between the constant capacitance
and triple layer models in that its "binding constants," KiA, can be related
in pairs to the intrinsic equilibrium constants of the constant capacitance
model whereas its "surface potentials," /JA' are used like the inner potentials in the triple layer model and assigned to specific planes containing
adsorbed species. Chemical reactions and their, concomitant equilibrium
constants play no role in the objective model. In this respect, the model is
close in spirit to the classical Gouy-Chapman-Stern-Grahame picture of
the electrical double layer.t" with the a plane identified as the inner
Helmholtz plane and the f3 plane as the outer Helmholtz plane. This
mixture of classical double layer theory and the surface complexation
concept is not a fully self-consistent chemical model, however, because it
does not reduce to the triple layer model when no species are present in the

a plane.

188

THE SURFACE CHEMISTRY OF SOILS

To see this point in detail, consider a hydrous oxide surface bathed by an


aqueous solution containing only the electrolyte ML(aq). In this case,
Eqs. 5.76 and Eq. 5.58 prescribe identical surface charge-inner potential
relationships (with G;'l= G~l + G;;rl = C1 and G(3d = C2 ) and Eqs. 5.77
and 5.59 are the same. The surface charge density in the s plane, according
to the objective model, is given by Eq. 5.80. This equation indicates that
O"H is determined by the activity of the proton in the aqueous solution
phase and by the inner potential, l/Js' On the other hand, according to the
triple layer model, O"H is given by Eq. 5.63, which, through Eqs. 5.60 and
5.61, depends explicitly on the activity of H+, M+, and L - in the aqueous
solution phase, as well as on the inner potentials l/Js and l/J(3' Indeed,
Eqs. 5.60a, 5.61a, and 5.79 are in direct conflict, since the quantity Qal
contains XSOH!, the mole fraction of uncomplexed SOH~, whereas
r~~Ht in Eq. 5.79 comprises the sum XSOHt + xSOH2L, the total mole
fraction of SOH~ , both complexed and uncomplexed. The root cause of
this difference is the inconsistent use of a f3 plane in the objective model to
allow adsorption of the background electrolyte without then partitioning
O"H to reflect this adsorption. Therefore, despite the very good description
of experimental data possible with the objective model,36,39 it does not
have the same status, as a chemical model, accorded the constant capacitance and triple layer models.
5.6. THE STRUCTURE OF SURFACE COMPLEXATION MODELS
The intent of the present chapter has been to elucidate the molecular
structural features of several models of the interfacial region that enjoy
widespread application. It should be apparent from the discussion that
these models are only qualitatively accurate at the molecular level, even
though good quantitative descriptions of potentiometric titration data and
ion adsorption isotherms can be obtained by curve-fitting techniques. The
conclusion that should be drawn from these facts is that titration and
adsorption experiments are not sensitive to the detailed structure of the
interfacial region. Given the typical number of adjustable parameters in a
surface complexation model, it is possible to fit most data on proton, metal
cation, and oxyanion adsorption without having an accurate picture of
surface speciation." More severe constraints on a model, however, might
be imposed by data obtained for aqueous systems comprising several
adsorptive ions involved in competitive surface reactions." but the most
direct sources of information concerning the' molecular structure of the
interfacial region are X-ray, optical, magnetic resonance, and neutron
spectroscopy experiments.F These experiments offer probes of molecular
energy transfer that are responsive to the details of structure and bonding
at an interface. They can be expected to be reliable guides in the future
development of surface chemical models.
It is possible to distill from the discussion in Sees. 5.1 to 5.5 a set of
general properties for surface complexation models. These properties

CHEMICAL MODELS OF SURFACE COMPLEXATION

189

divide naturally into two categories: constraint equations and molecular


hypotheses. An illustration of the two categories is now made with the
inorganic surface hydroxyl group taken as an example.
For each class of reactive surface hydroxyl groups,
there is an equation of mole balance. This equation has the general
mathematical form
CONSTRAINT EQUATIONS.

~S = L L L a{(S(i)O)aM;P(OH)yHxL~k)}
a j,k p,y,k,q

LL L
b

j,k p,y,x,q

b{S~)M;,n(OH)yHxL~k)}

(5.82)

where M, is the total number of reactive hydroxyl groups of type i, S(i)OH,


per unit area, S is a specific surface area, and { } refer to a solid-phase
concentration in moles per kilogram. The species in the first sum in
Eq. 5.82 include as special cases the inner-sphere proton and hydroxide
ion complexes in Eq. 5.41, the inner-sphere metal complexes in Eq. 5.50,
the outer-sphere metal complexes in Eq. 5.67, and the outer-sphere ligand
complex in Eq. 5.70. The species in the second sum in Eq. 5.82 are
generalizations of the inner-sphere ligand complex in Eq. 5.37b.
For each of the surface charge densities UH, UIS, and Uos, an equation
that includes the relevant surface complexes can be established. The
surface density of net proton charge is given by. the expression
UH

= (FjS)

~
I

[L {S(')(OH)xL~k)} ,L
-

k,q,x

j,p,y

{S(i)OMi/\OH)y}] (5.83)

where F is the Faraday constant. Equation 5.83 is the same as Eq. 5.46
with the one difference that here the species that contribute to XSOH! and
XSo- are shown explicitly. The surface density of inner-sphere complex
charge is given by the equation
UIS

= (FjS)

LLL

j,k p,y,x,q

[ZIs{(S(i)O)aM~j)(OH)yHxqk)}

ZIs{Sbi)M~j)(OH)yHxqk)}]

(5.84)

where, in the notation of Eq. 5.37, ZIS = pm + x - ql - y is the valence


of a complexed species bound directly to (S(i)O)a or S~). The first term in
the sum in Eq. 5.84 refers only to inner-sphere complexes with metal-like
species; the second term refers to inner-sphere complexes with ligand-like
species." The surface density of outer-sphere complex charge is given by
the equation
Uos = (FjS)

LL L
I

where Zos .. pm

+x

l,k

".y.x.t{

Zos{(S(i)O)aM~j)(OH)yHxL~k)}

(5.85)

- ql - y is the valence of a metal-like or ligand-like

190

THE SURFACE CHEMISTRY OF SOilS

species with at least one water molecule interposed between it and the
surface anion, (S(i)O)~-. Equations 5.83 to 5.85 are subject to the
charge-balance expression (Eq. 3.3b)
+ aH + aIS + aos + aD + 0
(5.86)
This equation and Eq. 5.82 provide the mass and charge balance constraints on any surface complexation model.
Since each of the solid phase species in Eq. 5.82 can be regarded as one
of the products in a chemical reaction, as in Eq. 5.37, there exist chemical
equilibrium constraints for each class of reactive surface hydroxyl groups:
S(i)O) M(j)(OH) H L(k(H)a-x
K(i) a
pyx q
(5 87 )
soc - (S(i)OH)"(M(j)P(OH)Y(L(kq
. a
ao

(S(i)M(j)(OH) H L(k(OH)b- y
b
Pyx q
sc (S(i)OH)b(M(j)p(HY(L(kq

K(') -

(5 87b)
.

where K~bc and K~b are thermodynamic equilibrium constants. The


reduction of the activities in Eq. 5.87 to products of activity coefficients
with either solid phase mole fractions or aqueous phase molalities requires
a choice of reference state. Either the Constant Ionic Medium Reference
State (as in the constant capacitance model) or the Infinite Dilution
Reference State (as in the triple layer model) may be chosen, and this
choice determines whether the ions in the background electrolyte are
included in Eqs. 5.82 to 5.85. Regardless of the choice of reference state,
the left sides of Eqs. 5.87 are independent of the composition of the solid
and aqueous phases and in this way act as constraints in addition to those of
mass and charge balance.
MOLECULAR HYPOTHESES.

If experimental methods exist for measuring the

composition of the solid and aqueous solution phases in a suspension of


adsorbent particles, then Eqs. 5.82 to 5.85 constitute a description of the
surface chemistry that requires no additional molecular hypotheses. 12
. Surface complexation models represent molecular theories that try to
calculate the composition of the solid phase in a suspension, given the
composition of the aqueous solution phase and a set of hypotheses
concerning the detailed structure of the interfacial region. The specific
focus of these models is the equilibrium constants in Eq. 5.87. Consider,
for example, Eq. 5.87a. The equilibrium constant K~bc can be expressed
(')

K(i) _ cK;oc fsoc


soc fa

(5.88)

SOH

where
CK(I)

soc -

xsodH)a
xS'OH(C)

(5.89)

is the conditional equilibrium constant, in the notation of Eq. 5.27, and

191

CHEMICAL MODElS OF SURFACE COMPLEXATION

fsoc and fSOH are rational activity coefficients.P Given that K~bc can be
measured, the task of a surface complexation model is to develop a method
for calculating the rational activity coefficients. Once this is accomplished,
the value of a: can be calculated and the composition of the solid phase
predicted through a prediction of Krbc and introduction of the results
into Eqs. 5.82 to 5.85. The prediction of the solid phase composition
proceeds in exactly the same way as a speciation calculation for the
aqueous solution phase."
In Sec. 5.2 it is shown that the van der Waals model in statistical
mechanics leads to a specific prediction of the functional dependence of
ratios like fsoc/HoH on the molecular properties of an interfacial region
(d. Eq. 5.36). The surface complexation models discussed in the present
chapter are related to the van der Waals model and therefore also make
specific predictions of the dependence of rational activity coefficients on
molecular parameters. As an example, consider the constant capacitance
model and, in particular, Eqs. 5.41b, 5.42b, and 5.43b. Equation 5.42b is a
special case of Eq. 5.89 (a = 1; P = Y = x = q = 0), and Eq. 5.43b is
analogous to Eq. 5.88. With the help of Eqs. 5.47 and 5.49b, one can write
the exact analog of Eq. 5.88:
C

K 2 = K~2 exp[F(-!o/go- - o/s)/RT]

(5.90)

since K 2 in Eq. 5.49b is a special case of K~~c in Eq. 5.87a. It follows from
a comparison of Eqs. 5.88 and 5.90 that

-fso-f:SOH

0
= exp [ F (120/S0-

/]
O/S )RT

(5.91)

in the constant capacitance model. Equation 5.91 is a fundamental molecular hypothesis in the model. Since the inner potential, o/s' cannot be
determined independently, the model must make the additional assumption contained in Eq. 5.39. Once this is done, only the capacitance
parameter C remains to be determined from measurements of the dependence of K~2 on up, as described in Sec. 5.3.
In the case of the triple layer model, the formation of the species SO- is
described by two equilibrium constants and Eqs. 5.60b and 5.60d are the
relevant special cases of Eq. 5.89. Equations 5.61, for i = 2, a = M+ are
analogous to Eq. 5.88. Taking Eq. 5.61b with a = M+ in combination
with Eq. 5.66, one can drive the expression

K = *Qm+ exp[ -e(-!o/goM

+ O/S ""'- o/(3)/kB T ]

(5.92)

since Kin Eq, 5.66 is a special case of K~bc in Eq. 5.87a. The ratio of
rational activity coefficients, fSOM/fsOH' then follows from a comparison of
Eqs. 5.88 and 5.92:
f:f:SOM
so...

= exp[

e(~o/~OM + o/M -

o/(3)/k BT]

(5.93)

This equation expresses a fundamental molecular hypothesis in the triple

Table 5.3. Constraint equations and molecular hypotheses in the constant capacitance and triple layer
models for a hydrous oxide suspended in a 1: 1 electrolyte solution.
Constraint equation

Molecular hypothesis

Constant capacitance model


= f~oc exp(ZsocFr/JslRT)
aH = Cr/Js for aH near zero
No outer-sphere surface complexes

XSOH{ + XSO- + XSOH = 1


aH ~ (FM/ N A)(XSOW2 - XSO-)
aH + aD = 0

fsoc

Triple layer model


xSOHi + XSOH2L + xso- + XSOM + XSOH = 1
aH = (FM/NA)(xsOHi + XSOH2L - xso- - XSOM)
aos = (FM/NA)(xsOM - XSOH2d
aH + ,aos + aD = 0

fil'

= f;~ exp(Z;er/JA/kBT)

(A

=s

or (3)

aH = C1(r/Js - r/Jf3)
aH + aos = C 2(r/Jf3 - r/Jd)
aD

-(8E ODRTC Md l / 2 sinh(Fr/Jd/2RT)

Note the following:


Zsoc = valence of the surface complex SOc.
M
= reactive hydroxyl groups per unit area.
IiA = rational activity coefficient of surface species i in plane A.
I~
= constant determined by the reference state assumed.
Z,
= valence of surface species i (e.g., SO-; M+ complexed by SO-; H+ complexed by SOH) located in the plane .1.(.1.
CM L = concentration of ML(aq).

= s or f3).

CHEMICAL MODELS OF SURFACE COMPLEXATION

193

layer model. Because the inner potentials are not determinable independently, Eqs. 5.58a and 5.64 must be invoked, leaving C1 to be determined
by measurements of the dependence of * QM+ on UH'
The general pattern that can be extracted from these two examples
comprises two steps:
1. A molecular interpretation of the equilibrium constants in Eq. 5.87 is
initiated by stating whether the surface species in the numerators form
inner-sphere or outer-sphere complexes.
2. Consistent with the hypotheses made in step 1, the rational activity
coefficients of the surface species are expressed mathematically in terms
of molecular parameters. If these parameters cannot be related directly
to measurable quantities, additional molecular equations are invoked to
provide such a relationship.

Thus the molecular hypotheses in a surface complexation model provide a


detailed interpretation, in terms of an assumed structure for the interfacial
region, of the concentrations, surface charge densities, and equilibrium
constants in the constraint equations. These features of the models are
illustrated in Table 5.3 for an adsorbent bearing a single class of surface
hydroxyl group and suspended in a 1: 1 background electrolyte solution.
Note that the expression given for the rational activity coefficient in the
triple layer model is consistent wth Eqs. 5.61, 5.69, and 5.71b when
applied to surface species, not surface complexes. Equation 5.93 results
from applying the expression in Table 5.3 to SO- and complexes M+
individually, then setting (f~o-sf~+{3) = exp( -e/JgoM/2kBT). The same
general considerations apply to an adsorbent bearing other kinds of surface
functional group, e.g., the siloxane ditrigonal cavity.t" It is evident from an
examination of Table 5.3 that the basis for improvement in the development of surface complexation models lies with more accurate specifications
of inner- and outer-sphere surface complexes and more comprehensive
expressions for the rational activity coefficients of surface species.
NOTES
1. Introductory discussions of the Poisson-Boltzmann equation can be found in
Sec. III of K. L. Babcock, Theory of the chemical properties of soil colloidal
systems at equilibrium, Hilgardia 34: 417 (1963); in Chap. 6 of G. Sposito, The
Thermodynamics of Soil Solutions (Clarendon Press, Oxford, 1981); and in
Chap. 1 of G. H. Bolt, Soil Chemistry. B: Physico-Chemical Models (Elsevier,
Amsterdam, 1979).
2. See, e.g., D. C. Grahame, The electrical double layer and the theory of
electrocapillarity, Chern. Rev. 41: 441 (1947); C. W. Outhwaite, Modified
Poisson-Boltzmann equatiorl in electric double layer theory based on the
Bogoliubov-Born-Green-Yvon integral equations, J. C. S. Faraday 1/74: 1214
(1978); and the references cited therein.
3. The situation in which this condition is not met (i.e .. interacting double layers)
is discussed in E. C. Childs. The splice charge in the Gouy layer between two

194

THE SURFACE CHEMISTRY OF SOILS

plane, parallel non-conducting particles, Trans. Faraday Soc. 50: 1356 (1954)
and in F.A.M. de Haan, The negative adsorption of anions (anion exclusion)
in systems with interacting double layers, J. Phys. Chern. 68: 1970 (1964). See
also Chap. 1 and 7 in G. H. Bolt, op. cit.' These more complicated DDL
theories are not required for the chemical models discussed in the present
chapter.
4. D. C. Grahame, Diffuse double layer theory for electrolytes of unsymmetrical
valence types, J. Chern. Phys. 21: 1054 (1953). See also Chap. 1 in G. H. Bolt,
. 1
op. elf.
5. See, e.g., G. H. Bolt, op cit. ,1 pp. 14-17: Most often, the absolute magnitude
of r/J(O) is set arbitrarily equal to + 00 and the DDL model is calibrated instead
by introducing (T(x) = -(Tin into Eq. 5.5, where (Tin is a measured value of the
intrinsic surface charge density. The adjustable parameter in the model is
thereby shifted from r/J(O) to the value x = 5 that satisfies Eq. 5.5, with an
appropriate formula for r/J(x) used in calculating the derivative on the left side.
The effect of this substitute calibration is to place the plane to which (Tin refers
at some positive value of x instead of at x = 0 and to restrict the application of
the DDL model to the region x > 5. However, aside from the advantage
gained by forcing (T(x) to have a reasonable value at the particle surface, the
physical aspects of this renormalized DDL model vis-a-vis surface complexes
are the same as those of the unrenormalized model having the surface plane at
x = O.
6. The first unequivocal demonstration of this important, well-known constraint
appears in O. Stern, Zur Theorie der elektrolytischen Doppelschicht,
Z. Elektrochem. 30: 508 (1924). The Stern model of the interfacial region was
the first chemical model in the spirit of the present chapter.
7. G. M. Torrie and J. P. Valleau, Electrical double layers. 1: Monte Carlo study
of a uniformly charged surface, J. Chern. Phys. 73: 5807 (1980).4. Limitations
of the Gouy-Chapman theory, J. Phys. Chern. 86: 3251 (1982). The second
paper also gives references for recent attempts to modify DDL theory to take
into account finite ion size and potential fluctuations.
8. Theoretical calculations that anticipated this result are summarized in Chap. 7
of M. J. Sparnaay, The Electrical Double Layer. Pergamon Press, Oxford,
1972.
9. The development of computer-based algorithms from surface complexation
models is discussed in J. Westall, Chemical equilibrium including adsorption
on charged surfaces, in Particulates in Water (M. C. Kavanaugh and J. O.
Leckie, eds.). American Chemical Society; Washington, D.C., 1980.
10. An introduction to statistical mechanics that is adequate for a comprehension
of the present discussion is given in Chap. 6 of G. Sposito, op cit.: Full details
are provided in Chap. 7 and 14 of T. L. Hill, An fntroduction to Statistical
Thermodynamics. Addison-Wesley, Reading, Mass., 1960.
11. The method of deriving Eq. 5.23 presented here follows the discussion on
p. 252 in T. L. Hill, op. cit. 10 More general (and more rigorous) methods for
deriving partition functions in the van der Waals limit are given in Sec. 16 of
E. A. Guggenheim, Statistical thermodynamics of mixtures with nonzero
energies of mixing, Proc. Royal Soc. (London) 183A:213 (1944).
12. G. Sposito, On the surface complexation model of the oxide-aqueous solution
interface, J. Col/oid Interface Sci. 91:329 (1983).
13. Conditional equilibrium constants for reactions in heterogeneous chemical
systems are discussed fully in Chap. 5 of O. Sposito, op. cit. I

CHEMICAL MODELS OF SURFACE COMPLEXATION

195

14. W. Stumm, C. P. Huang, and S. R. Jenkins, Specific chemical interaction


affecting the stability of dispersed systems, Croatica Chern. Acta 42: 223 (1970).

15.

16.
17.
18.

19.

20.

21.

22.

23.

W. Stumm, H. Hohl, and F. Dalang, Interaction of metal ions with hydrous


oxide surfaces, Croatica Chern. Acta 48:491 (1976). W. Stumm, R. Kummert,
and L. Sigg, A ligand exchange model for the adsorption of inorganic and
organic ligands at hydrous oxide interfaces, Croatica Chern. Acta 53: 291
(1980).
Comprehensive recent discussions of the constant capacitance model are given
in P. W. Schindler, Surface complexes at oxide-water interfaces, in Adsorption
of Inorganics at Solid-Liquid Interfaces (M. A. Anderson and A. J. Rubin,
eds.; Ann Arbor Science, Ann Arbor, Michigan, 1981) and in H. Hohl, L.
Sigg, and W. Stumm, Characterization of surface chemical properties of oxides
in natural waters, in M. C. Kavanaugh and J. 0. Leckie, op. cit.9 A computerbased algorithm for the constant capacitance model is described by J. Westall,
. 9
op. Cit.
The Constant Ionic Medium Reference State is compared with the more
common Infinite Dilution Reference State in Chap. 2 of G. Sposito, op. cit."
H. Hohl and W. Stumm, Interaction of Pb2+ with hydrous y-Al z0 3 , J. Colloid
Interface Sci. 55: 281 (1976).
The standard state to which conventional thermodynamic equilibrium constants refer is that of unit mole fraction for each solid phase component in a
reaction and unit molality without coulomb interactions for each aqueous
phase ion in a reaction. See p. 32 in G. Sposito, op. cir?
S.-Y. Chu and G. Sposito, The thermodynamics of ternary cation exchange
and the subregular model, Soil Sci. Soc. Am. J. 45: 1084 (1981). The discussion
given on p. 1088 of this paper, particularly Eq. 35, shows that the van der
Waals potentials Zjel/Jp must be independent of the index i for surface
complexes of a given kind (e.g., inner-sphere).
c.-P. Huang and W. Stumm, Specific adsorption of cations on hydrous
y-Alz0 3 , J. Colloid Interface Sci. 43:409 (1973). P. W. Schindler, B. Furst,
R. Dick, and P. U. Wolf, Ligand properties of surface silanol groups. I.
Surface complex formation with Fe3+, Cu z+ , Cd z+ , and Pb2+, J. Colloid
Interface Sci. 55:469 (1976). C.-P. Huang and Y. T. Lin, Specific adsorption of
Co(II) and [Co(III)EDTA]- complexes on hydrous oxide surfaces, in Adsorptionfrom Aqueous Solution (P. H. Tewari, ed.). Plenum, New York, 1981.
A similar interpretation of the data in Fig. 4.7 is given in D. W. Fuerstenau,
D. Manmohan, and S. Raghavan, the adsorption of alkaline-earth metal ions
at the rutile/aqueous solution interface, in P. H. Tewari, op. cit. zO
R. Kummert and W. Stumm, The surface complexation of organic acids on
hydrous y-Al z0 3 , J. Colloid Interface Sci. 75: 373 (1980). L. Sigg and
W. Stumm, The interaction of anions and weak acids with the hydrous goethite
(a-FeOOH) surface, Colloids Surfaces 2: 101 (1980-1981).
D. E. Yates, S. Levine, and T. W. Healy, Site-binding model of the electrical
double layer at the oxide/water interface, J.es. Faraday 170: 1807 (1974).
J. A. Davis, R. 0. James, and J. 0. Leckie, Surface ionization and complexation at the oxide/water interface. I: Computation of electrical double layer
properties in simple electrolytes, J. Colloid Interface Sci. 63: 480 (1978). J. A.
Davis and J. 0. Leckie, Surface ionization and complexation at the
oxide/water interface. II: Surface properties of amorphous iron oxyhydroxide
and adsorption of metal ions, J. Colloid Interface Sci. 67: 90 (1978). 3:
Adsorption of IInionl, J. Colloid Interfufe Sci. 74: 32 (1980).

196

THE SURFACE CHEMISTRY OF SOilS

24. Comprehensive recent surveys of the triple layer model are to be found in J. A.
Davis and J. O. Leckie, Speciation of adsorbed ions at the oxide/water
interface, in Chemical Modeling in Aqueous Systems (E. A. Jenne, ed.;
American Chemical Society, Washington, D.C., 1979) and in R. O. James and
G. A. Parks, Characterization of aqueous colloids by their electrical doublelayer and intrinsic surface chemical properties, Surface Colloid Sci. 12: 119
(1982). A computer-based algorithm for the triple layer model is described by
J. Westall, op. cit. 9
25. R. O. James, J. A. Davis, and J. O. Leckie, Computer simulation of the
conductometric and potentiometric titrations of the surface groups on ionizable
latexes, J. Colloid Interface Sci. 65: 331 (1978). A complete discussion of the
graphic extrapolation methods, with examples, is given on pp. 167-170 in
R. O. James and G. A. Parks, op. cit. 24
26. J. A. Davis, R. O. James, and J. O. Leckie, op. cit.23
27. R. O. James, P. J. Stiglich, and T. W. Healy, The Ti0 2/aqueous electrolyte
system: Applications of colloid models and model colloids, in P. H. Tewari,
op. cit.2o
28. L. Balistrieri and J. W. Murray, Surface of goethite (a-FeOOH) in seawater,
in E. A. Jenne, op. cit. 24 The surface chemistry of goethite (a-FeOOH) in
major ion seawater, Am. J. Sci. 281: 788 (1981). The adsorption of Cu, Pb,
Zn, and Cd on goethite from major ion seawater, Geochim. Cosmochim. Acta
46: 1253 (1982). The surface chemistry of 8-Mn02 in major ion seawater,
Geochim. Cosmochim. Acta 46: 1041 (1982).
29. A. E. Regazzoni, M. A. Blesa, and A. J. G. Maroto, Interfacial properties of
zirconium dioxide and magnetite in water, J. Colloid Interface Sci. 91: 560
(1982).
30. M. M. Benjamin and N. S. Bloom, Effects of strong binding of anionic
adsorbates on adsorption of trace metals on amorphous iron oxyhydroxide, in
20
PHT
. . ewan,' op. cit.
31. Surface speciation calculations are analogous to speciation calculations for
aqueous solution phases, as pointed out by J. Westall, op. cit. ,9 in a detailed
review.
32. See Part II of the series by J. A. Davis and J. O. Leckie, op. cit. 23
33. Other examples of surface chemical data that are described equally well by the
triple layer and constant capacitance models are given in J. Westall and
H. Hohl, A comparison of electrostatic models for the oxide/solution interface, Advan. Colloid Interface Sci. 12: 265 (1980).
34. See R. O. James et aI., op. cit. 27
35. See Part 3 of the series by J. A. Davis and J. O. Leckie, op. cit. 23 ; J. A. Davis
and J. O. Leckie, op. cit.24; L. Balistrieri and J. W. Murray, op. cit28
36. M. Sasaki, M. Morlya, T. Yasunaga, and R. D. Astumian, A kinetic study of
ion-pair formation on the surface of a-FeOOH in aqueous suspension using the
electric field pulse technique, J. Phys. Chem. 87: 1449 (1983).
37. J. W. Bowden, A. M. Posner, and J. P. Quirk, Ionic adsorption on variable
charge mineral surfaces: Theoretical-charge development and titration curves,
Aust. J. Soil Res. 15: 121 (1977). N. J. Barrow, J. W. Bowden, A. M. Posner,
and J. P. Quirk, An objective method for fitting models of ion adsorption on
variable charge surfaces, Aust. J. Soil Res. 18:34 (1980). A. M. Posner and
N. J. Barrow. Simplification of a model for ion adsorption on oxide surfaces,
J. Soil Sci. 33: 211 (19H2). Reviews of the objective model are given in

CHEMICAL MODELS OF SURFACE COMPLEXATION

38.
39.

40.
41.

42.
43.
44.

197

F. J. Hingston, A review of anion adsorption, in M. A. Anderson and A. J.


Rubin, op. cit.,15 and in J. W. Bowden, A. M. Posner, and J. P. Quirk,
Adsorption and charging phenomena in variable charge soils, in Soils with
Variable Charge (B.K.G. Theng, ed.; New Zealand Society of Soil Science,
Lower Hutt, New Zealand, 1980). The objective model is compared briefly
with the constant capacitance and triple layer models by J. Westall and
H. Hohl, op. cit. 33
See, e.g., Eq. 7-8 in T. L. Hill, op. cit. lO
J. W. Bowden, S. Nagarajah, N. J. Barrow, A. M. Posner, and J. P. Quirk,
Describing the adsorption of phosphate, citrate and selenite on a variablecharge mineral surface, Aust. J. Soil Res. 18: 49 (1980). N. J. Barrow,
J. W. Bowden, A. M. Posner, and J. P. Quirk, Describing the adsorption of
copper, zinc and lead on a variable charge mineral surface, Aust. J. Soil Res.
19: 309 (1981).
See, e.g., D. C. Grahame, op. cit.,2 for a discussion of the classical electrical
double layer.
This central point is made cogently by J. Westall and H. Hohl, op. cit.33 Even
the diffuse double layer model can be made to fit adsorption data by using the
value aD as an adjustable parameter!
See, e.g., J. W. Stucki and W. L. Banwart, Advanced Chemical Methods for
Soil and Clay Minerals Research. D. Reidel, Dordrecht, Holland, 1980.
A complexed species is metal-like if ZIS > 0, ligand-like if ZIS < O.
The constant capacitance model has been applied, in essence, to the adsorption
of monovalent cations by montmorillonite in 1. Shainberg and W. D. Kemper,
Ion exchange equilibria on montmorillonite, Soil Sci. 103: 4 (1967). The triple
layer model is applied to the same phenomenon by R. O. James and G. A.
. 24
Par k s, op. elf.

FOR FURTHER READING

M. A. Anderson and A. J. Rubin, Adsorption of Inorganics at Solid-Liquid


Interfaces. Ann Arbor Science, Ann Arbor, Michigan, 1981. Chapters 1, 2, 5, 6 and
7 provide specialized introductions to surface complexation models.
G; H. Bolt, Soil Chemistry. B: Physico-Chemical Models, 2nd rev. ed. Elsevier,
Amsterdam, 1982. Chapter 2 surveys diffuse double layer theory as applied in soil
chemistry, and Chap. 13 surveys surface complexation models.
R. O. James and G. A. Parks, Characterization of aqueous colloids by their
electrical double-layer and intrinsic surface chemical properties, Surface and
Colloid Science 12: 119 (1982). Perhaps the most complete review of the triple layer
model from the perspective of Gouy-Chapman-Stern-Grahame double layer
theory.
M. C. Kavanaugh and J. O. Leckie, Particulates in Water: Characterization, Fate,
Effect, and Removal. Advances in Chemistry Series 189. American Chemical
Society, Washington, D.C. 1980. Chapters 1 and 2 give excellent accounts of the
constant capacitance model in theory and application. Chapter 2 also contains a
discussion of the computational aspects of the triple layer model.

6
SURFACE CHEMICAL ASPECTS
OF SOIL COLLOIDAL STABILITY

6.1. THE SMECTITE QUASICRYSTAL

In Sec. 1.4 the concept of the smectite quasicrystal was introduced as a


means of interpreting data on the specific surface area of montmorillonite.
This concept, which describes the heterogeneous smectite particle formed
by stacking hydrated unit layers along the crystallographic c axis
(Fig. 1.11),1 is discussed in more detail in the present section. The
discussion is intended to fill out the conceptual basis for understanding the
behavior of soil clay suspensions containing smectites. Fundamentally, for
any colloidal suspension, an operational meaning must be given to the
terms particle and interparticle forces in order to describe the structure and
stability of the suspension. The present section addresses the nature of the
particles in smectite suspensions, taking as an illustrative example the clay
mineral montmorillonite.
Perhaps the first clear indication of the existence of montmorillonite quasicrystals appeared 30 years ago with the
finding that Ca-montmorillonite suspensions exhibit a d(001) spacing of
1.91 nm that persists even as the electrolyte concentration is reduced to
zero.f This invariant d(001) spacing is consistent with the formation of an
outer-sphere surface complex between exchangeable Ca2+ ions and a pair
of opposing siloxane ditrigonal cavities. In this surface complex, the
octahedral solvation unit, Ca(H20)~+, would be arranged in the interlayer
region with its principal symmetry axis parallel to the crystallographic
c-axis. Four of the solvating water molecules then would lie in a central
plane that is parallel with the opposing siloxane surfaces, while the
remaining two water molecules would reside in planes between the
siloxane surfaces and the central plane (see Fig. 6.1). Supporting evidence
for this structural arrangement can be found in the results of quasielastic
neutron scattering experiments. which point to the existence of a rigid
EQUILIBRIUM STRUCTURE.

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

199

Figure 6.1. An exploded view of the three-layer hydrate of Ca-montmorillonite.

octahedral solvation shell for Ca2+ in the 1.91 nm hydrate of Camontmorillonite, and in electron spin resonance spectra of exchangeable
Cu2+ and Mn2 + in the three-layer hydrate of montmorillonite, which show
that the principal symmetry axis of the octahedral solvation shell is
perpendicular to the siloxane surface." Thus it may be concluded that a
stable, outer-sphere surface complex that brings together two unit phyllosilicate layers in face-to-face association is a characteristic molecular
structure in suspensions of montmorillonite bearing bivalent exchangeable
cations.
A variety of experimental methods has been applied to determine the
equilibrium number of unit layers in a montmorillonite quasicrystal. The
results of some of these experiments are summarized in Table 6.1.
Although variability exists among the experimental estimates for a given
homoionic form of the clay mineral, a clear consensus emerges as to the
absence of quasicrystals in dilute Na-montmorillonite suspensions
(1.2 0.2 unit layers) and to the presence offour to eight unit layers per
quasicrystal in dilute Ca-montmorillonite suspensions. Perhaps none of the
determinations summarized in Table 6.1 is conclusive by itself, but taken as
a group the set of data is quite convincing. A brief review of the five
techniques listed in Table 6.1 can serve as a guide to estimating the
reliability of the individual results.
Small-angle neutron or X-ray scattering" from dilute (less than 20 kg of
clay per cubic meter) suspensions of montmorillonite can be used to infer
the average thickness of the scattering particles, which are assumed to be
disc-shaped aggregates consisting of unit layers in face-to-face association
with water layers interspersed between them. The intensity of the neutrons
or X-rays scattered through small angles by one of these aggregates
depends on the radius of gyration, which in turn can be related to the

Table 6.1. Experimental estimates of the number of unit layers in face-to-face association in a particle of
montmorillonite
Exchangeable
cation

Neutron
scattering"

Li+
Na+
K+
Cs+
Mg 2+
Ca 2+

1
2
3
-

X-ray
scattering"

Electron
microscopy'
-

1
-

6-8

1.4 0.5
-

16 2

Viscosity"

1.0

1.0 1.3
1.4
1.7
2.9
4.2 5.0
-

Light
scattering7

Chloride
exclusions

1.0 1.0 1.0


1.2 1.7 1.2
1.5 2.7 2.8 3.0 5.5 4.2 6.2 7.0 2.7

1.0 1.0
1.2 1.1
1.5 2.1
4.2
8.8
5.9 6.5

SURFACE CHEMICAL ASPECTS OF SOil COllOIDAL STABILITY

201

particle thickness through standard geometric formulas. Small-angle scattering data interpreted in this way for cold neutrons led to aggregate
thickness estimates of 1.1, 2.6, and 4.2 nm, respectively, for Li-, K-, and
Cs-montmorillonite." Within the experimental precision of 0.2 nm, these
results are consistent with particle structures comprising one, two, and
three unit phyllosilicate layers, respectively (e.g., in K-montmorillonite,
2.6 0.2 nm compared with 0.96 + 0.29 + 0.96 = 2.2 nm for a two-layer
quasicrystal). The same kind of approach with X-rays led to estimates of
one unit layer per particle in dilute suspensions of Na-montmorillonite and
six to eight unit layers per particle for Ca-montmorillonite."
High-resolution transmission electron microscopy' can be used to make a
direct estimate of the number of unit layers in a montmorillonite aggregate. A critical factor in this method is sample preparation. If the
solidification and ultimate dehydration of a clay suspension (required in
order to present the suspension to an electron beam) do not alter the
aggregate structure, then the average number of layers in an aggregate
can be determined accurately through a microscopic examination of its
cross section. In practice, some alteration in aggregate structure is expected during sample preparation, and there is a possibility that ultrasectioning the sample will lead to an overestimate of the number of layers
in thick quasicrystals whose outermost clay platelets are only partially
intact." For these reasons, the estimate for Ca-montmorillonite in
Table 6.1 should be regarded as an upper limit.
Measurements of the specific viscosity of a very dilute (less than 2 kg of
clay per cubic meter) montmorillonite suspension can be used to determine
the average thickness of the suspended clay particles if the relationship"
71sp =

k 1Jp

(6.1)

is observed to represent the specific viscosity, 71sp, as a function of the


volume fraction of clay particles, 1Jp In Eq. 6.1,
71sp

71su -

71so

(6.2)

71so

and

1Jp = (c/Pc)(l + 0)

(6.3)

where 71su is the viscosity of a suspension containing c kilograms of clay per


cubic meter of suspension volume, 71so is the viscosity of the aqueous
solution in which the clay is suspended, Pc is the mass density (in kg-rn P)
of the clay, and 0 is the ratio of occluded water volume to clay volume in an
aggregate. If it is assumed that the aggregates are thin disc-ellipsoids of
semimajor axis a and semiminor axis b (b ~ a), then kin Eq. 6.1 can be
interpreted geometrically:"
k = 0.849a
h

(6.4)

202

THE SURFACE CHEMISTRY OF SOILS

The application of Eq. 6.1 to the estimation of the number of unit layers
per clay aggregate involves the additional assumptions that (1) the parameters a, fJ, and Pc do not depend on the type of exchangeable cation on
the clay and (2) each aggregate of either Na- or Li-montmorillonite
comprises a single unit layer. With these two assumptions, the slopes of
linear plots of 1]sp against c can be compared for several homoionic
montmorillonites and used to calculate the number of unit layers, as given
in Table 6.1. These estimates are relative values based on a number of
simplifying assumptions.
The use of light scattering to estimate the number of unit layers per clay
aggregate in a montmorillonite suspension is analogous to the method of
small-angle neutron or X-ray scattering. The intensity of light scattered by
an aggregate depends on the refractive index characteristics of the aggregate, the concentration of clay in the suspension, and the number of unit
layers in face-to-face association within the aggregate." Under the assumption that the refractive index characteristics and the mass density of the
clay do not depend on the type of exchangeable cation, the turbidity of a
series of homoionic montmorillonites is proportional to the number of unit
layers per aggregate. With the convention that Li-montmorillonite suspensions contain only single-layer aggregates, relative estimates, as given in
Table 6.1, are possible.
Graphs of the volume of chloride exclusion against the inverse square
root of chloride concentration can be used to estimate the specific surface
area of the aggregates in a montmorillonite suspension, according to
Eqs. 1.17 and 1.19. If it is assumed that the entire suspension comprises
quasicrystals with n unit layers each, thenf = 1.0 in Eqs. 1.19 and 1.20. If
it is assumed also that no complexes form between exchangeable cations
and siloxane ditrigonal cavities on the external surfaces-of quasicrystals,
then Eq. 1.20 leads to the conclusion that the specific surface area
measured by chloride exclusion is proportional to n, With the convention
that n = 1 for Li-montmorillonite, the relative estimates of n in the last
two columns of Table 6.1 can be made." These estimates are naive in that
surface complexes on the external surfaces are ignored despite the
evidence for their existence discussed following Table 1.8. However,
measurements of Vex for Ca- and Mg-montmorillonite suspensions do
indicate a limiting value, equal to about 0.3 dnr' kg-I, as the inverse
square root of the chloride concentration is extrapolated to zero." This
limiting value is in agreement with the y intercept in Eq. 1.19:
Sod/2 = 7.51 x 105 m2 kg- 1 x (1.91 - 0.96)nm/2 = 0.357 dnr' kg-I,
the volume inside the quasicrystals from which chloride is assumed to be
completely excluded. Although the estimates of n in Table 6.1 also reflect
differences in the extent of external surface complexation on the
homoionic montmorillonites, the good agreement of the measured Vex
values with those predicted by Eq. 1.19 indicates that chloride exclusion
data do give evidence for quasicrystal formation.

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

203

The estimates in Table 6.1 clearly are not truly quantitative since they
are based on both model assumptions and computational simplifications
that cannot be substantiated a priori. Nonetheless, they do show unanimity
for a given homoionic form of montmorillonite and it does seem justifiable
to infer from them that, in relatively dilute suspensions, the structure of
Na-montmorillonite particles is different from that of Ca-rnontmorillonite
particles. By induction one would conclude also that there should be some
kind of continuous transition from more or less single-unit-layer particles
to quasicrystals as one observes changes in the properties of suspensions of
montmorillonite bearing both Na + and Ca2+ on the basal planes. This
conclusion is verified experimentally by the data in Fig. 6.2.10 The ordinate
of the graph represents the ratio of the value of some suspension property
P to the value of P for Na-montmorillonite; the abscissa refers to values of
the charge fraction of exchangeable Na+ on the clay. Both the intensity' of
the light transmitted (not scattered) by a suspension and the electrophoretic mobility of its constituent particles are expected to be directly sensitive to particle dimensions. The intrinsic viscosity (the ratio of T/sp to c in
Eqs. 6.1 and 6.3) and the chloride exclusion volume are dependent on

Figure 6.2. Scaled value P/ PN a of four montmorillonite suspension properties


versus the charge fraction on Na+ on the clay.lOThe curves through the data points
are meant only as guides to the eye.
1.0

r-------:::~.--I

0.8

0.6

TRANSMITTED LIGHT
INTENSITY

ELECTROPHORETIC MOBILITY

INTRINSIC VISCOSITY

CHLORIDE EXCLUSION
VOLUME

oOl----...-.--~--......L_--L.---~
0.2

0.4

0.6

0.8

1.0

204

THE SURFACE CHEMISTRY OF SOILS

particle shape and external surface area, respectively. That these four
properties show sharp increases when the charge fraction of Na + on the
clay lies between 0.15 and 0.30 indicates that, in this range, the quasicrystals of Ca-montmorillonite are largely being broken up in favor of
more or less single-unit-layer aggregates. Evidently, when E N a < 0.3, the
quasicrystals are stable entities with exchangeable Na+ cations residing
principally on their external surfaces, as discussed in connection with
Fig. 3.4.
The mixing together of Na- and Camontmorillonite suspensions to produce an overall charge fraction of Na +
on the clay particles below 0.1 results in a very rapid (less than 1 min)
formation of quasicrystals from conversion of the Na-montmorillonite
particles.!' This rapid conversion is necessarily mediated by a redistribu-:
tion of the exchangeable cations such that Na + ions are relocated, as
required, to the external surfaces of already-formed quasicrystals that
contain Ca2+ ions on their internal surfaces. The relocation probably
involves replacement by Na+ of Ca2+ already on external surfaces since
the latter ions are likely to have a higher mobility than Ca2+ adsorbed
inside a quasicrystal. 12
On the other hand, if the mixing together of two homoionic montmorillonite suspensions produces an overall charge fraction of Na + larger
than 0.6, then there is a gradual decomposition of the initial set of
quasicrystals, with the final result being a suspension whose physical
properties resemble closely those of a collection of single-unit-layer clay
particles.!' This slower process of quasicrystal conversion to single unit
layers reflects both the lower mobility of Ca2+ in the interlayer region and
the difficulty Na+ has in replacing these bivalent cations in the initial step
of quasicrystal destruction. In keeping with this point of view, the decomposition is aided by maintaining a low electrolyte concentration, since
interlayer swelling is thereby enhanced. 13
These characteristics of quasicrystal formation and breakdown are
consistent with the trends in Fig. 6.2. When E N a is less than 0.1, for
example, the properties of a mixed Na/Ca-montmorillonite do not differ
much from those of Ca-montmorillonite and the quasicrystal should
remain a stable entity. When E N a is larger than 0.6, however, the mixed
Na/Ca-montmorillonite exhibits properties that are indistinguishable from
those of a Na-montmorillonite and a quasicrystal should be an inherently
unstable structural unit.
When Na- and Ca-montmorillonite are mixed together and quasicrystal
formation is favored, it is likely that the thermodynamic driving force is a
negative enthalpy change (heat release). Since the quasicrystal represents a
more ordered local arrangement of unit montmorillonite layers than what
exists in a dilute suspension of Na-montmorillonite , the entropy change
accompanying quasicrystal formation should be negative. This result in
turn means that the entropy contribution to the Gibbs energy change for
QUASICRYSTAL

FORMATION.

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

205

quasicrystal formation is positive and that the required negative Gibbs


energy change (because the quasicrystal is stable) must derive from a
negative enthalpy change. This enthalpy change can be measured by
comparing the heat of immersion from mixing Na- and Ca-montmorillonite
in water to produce a given value of E N a with the heat of immersion of an
Na/Ca-montmorillonite with the same charge fraction E N a already on its
surfaces.!" The heat of immersion that accompanies the mixing of the two
homoionic clays is the larger by about 5.9 kJ . mol,", which means that the
enthalpy change for quasicrystal formation is about - 5.9 kJ mol,". This
enthalpy change can also be measured by subtracting the enthalpy of
Na-Ca exchange (+ 1.7 0.2 kJ -molj ') from the total enthalpy change
observed when a CaCl z solution is added to a suspension of Namontmorillonite. IS The result is -4.9 0.2 kJ . mol.,", which is in reasonable agreement with the first estimate. That both values are negative
substantiates the conjectured thermodynamic driving force for quasicrystal
formation.

6.2. INTERPARTICLE FORCES IN PHYLLOSILICATE SUSPENSIONS


The particles in a stable phyllosilicate suspension can be envisioned, in an
ideal geometric sense, as comprising one or more unit layers stacked along
the crystallographic c axis (Sec. 1.4 and Fig. 1.11). If the stacking is not
extensive, the particles resemble thin wafers and it is reasonable to imagine
the forces they exert on one another as emanating from planar surfaces.
The nature of these interparticle forces is discussed in the present section
after this geometric simplification is adopted. Actual interparticle forces no
doubt are more complicated than the model interactions described below,
but the essential physical features of the true forces should be represented
adequately.
If the net charge on a phyllosilicate particle
is not zero, an electric field and a nonvanishing net volumetric charge
density are produced in the aqueous solution near the surface of the
particle. These two effects combine to produce an electric force that in turn
can act on a neighboring charged clay particle to accelerate it, unless the
particle is restrained by an external force. If the two charged particles and
the aqueous solution between them are quiescent, the restraining external
force is manifest simply as a pressure gradient that balances the net electric
force in a unit element of volume anywhere in the aqueous solution phase.
Consider two identical opposing planes arranged geometrically as in
Fig. 3.3, with the origin of coordinates set in one of them and with the
other located at the point x = d (replacing the plane of shear) along an
outward normal from the first. Each plane is to represent the surface to
which up in Eq. 3.2 refers. Within the aqueous solution between the
planes, the volumetric charge density of species i is, according to the
diffuse double layer theory. equal to the ith term on the right side of
THE ELECTROSTATIC FORCE.

206

THE SURFACE CHEMISTRY OF SOILS

Eq. 5.9. In differential form this charge density obeys the equation

RT
zpdpj

-cjZjFexp(-ZjFl/J/RT)dl/J

-pjdl/J

(6.5)

where pj is the charge density at some point x and l/J is the corresponding
inner potential. Upon summing both sides of Eq. 6.5 over all species and
making use of the well-known relation for the osmotic pressure of an ideal
electrolyte solution,

RT}2 (pJZjF)

(6.6)

one can derive the force-balance expression.'"

dP + pdl/J

(6.7a)

= 0

where p is given by Eq. 5.9. Since l/J(x) is related to the volumetric charge
density, p(x), through the Poisson equation (Eq. 3.26), Eq. 6.7a can be
transformed, with the help of Eqs. 3.26 and 5.4b, to the expression
1
(dl/J)
-d [ P(x) - -eoD
dx
2
dx

2] = 0

(6.7b)

where eo is the permittivity of vacuum and D is the dielectric constant of


the water between the planes. Equation 6.7b is equivalent to the equation

P(x) - 2"1 eoD (dl/J)2


dx

(6.7c)

where Pm = P(d/2) is the pressure at the point x = d/2, halfway between


the two planes. Equation 6.7c includes the condition that the net electric
field, -dl/J/dx, be zero at this halfway point.
A practical application of Eq. 6.7c cannot be made until the constant of
integration Pm is expressed in terms of molecular properties. This can be
done by rewriting Eq. 6.7a as the integral expression
pm

dP

= -

Po

fc"'m

p(l/J)dl/J

(6.8)

where Po is the pressure in the aqueous solution far from the two planes,
where l/J(x) may be set equal to 0, and l/Jm == l/J(d/2). Equation 5.9 can be
introduced into Eq. 6.8 to yield the result

r;

= Po

+ RT}2 cj[exp(-ZjFl/Jm/RT) - 1]

(6.9)

where all of the symbols in the


following Eqs. 1.11 and 5.1.
pressure required to maintain
density O'p a distance d apart

second term on the right side are as defined


This model expression gives the applied
the two planar surfaces bearing the charge
at equilibrium.

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

207

The full specification of either P(x) or Pm as a function of position


between the planes cannot be made without solving Eq. 5.1 subject to the
conditions
dl/J)
( -dx x=d/2 -0

l/J(d/2) = l/Jm

(6.10)

A first integration of Eq. 5.1 under these conditions produces an equation


analogous to Eq. 5.8:
dl/J
{2RT
dx = -sgn(l/J) 8 D ~ cj[exp( -ZjFl/J(x)/RT) - exp( -ZjFl/Jm/RT)]

}1/2

(6.11)
Equation 6.11, like Eq. 5.8, can be solved analytically for a single symmetric electrolyte that dissociates completely. 17 However, the result is only a
formal expression in terms of elliptic integrals that must be evaluated from
mathematical tables." Alternatively, one can transform Eq. 6.11 into an
analog of Eq. 5.11a:
dy
dx

= -sgn(Y)K[2(cosh

Zy - cosh ZYm)]1/2

(6.12)

and use an iterative, computer-base algorithm to solve it numerically. 18


(In Eq. 6.12, y = Fl/J/RT and K = (3co, as in Table 5.1.) With either
approach, Eq. 6.12 is integrated between x = 0 and x = d/2 to produce a
relationship between l/Jm' l/J(O) , and the interplane separation d. The
integration can be done with l/J(O), the particle surface potential, fixed at a
predetermined value, in which case u(O) follows from Eqs. 5.5 and 6.12:

u(O) = -sgn(yo)

8 0 DRTK

[2(cosh ZYo - cosh ZYm)]I/2

(6.13)

where Yo = Fl/J(O)/RT. Since u(O) = UD in the DDL model and UD = -up


by Eq. 3.3a, the condition of fixed l/J(O) implies that up depends on l/Jm and
d. If the integration of Eq. 6.12 is done with up = -u(O) fixed instead,
then the boundary condition
(6.14)
must be applied and Eq. 6.13 is used to calculate l/J(O), which clearly
depends on l/Jm and d. It is possible that neither up nor l/J(O) should be held
fixed in the integration of Eq. 6.12 if surface complexation phenomena are
important as the interplane separation is varied.l" In this case, Eqs. 6.13
and 6.14 collapse into a single expression (because of Eqs. 3.3a and 5.5)
that relates a variable surface charge density to a variable surface potential.
Another relationship between these two parameters is needed in order to
make the algebraic problem determinate. This relationship can be taken
from any of the surface.complexation models described in Chap. S. For

208

THE SURFACE CHEMISTRY OF SOILS

example, Eqs, 5.42, 5.43, 5.44, and 5.46 can be used to connect O"p with
1/1(0), given that measured values of O"max and the K~(int) (i = 1,2) are
available.
'
The mathematical effort to solve Eq. 6.12 is reduced dramatically if the
condition ZKdj2 ~ 1 obtains. In this case, the approximate equation for
I/I(x) in Eq. 5.14 can be applied immediately to calculate I/Im as the
superposition of I/I(dj2) for each charged planer'?

RT

I/Im = 8 ZF a exp( - ZKd/2)

(6.15)

where a = tanh (ZFI/I(0)/4R T). Since ZKd/2 is very large, I/Im is very small
in this approximation and Eq. 6.9 must be expanded in a MacLaurin series
in I/Im to the lowest nonvanishing order to be consistent:
Pm = Po + cRT(ZFl/lm/RT)2
= Po + 64a 2cRT exp( -ZKd)

(6.16)

In this approximation, Pm is an exponentially decreasing function of the


interplane separation. The potenial energy per unit area, 'Pm' corresponding to the force per unit area, Pm' can be calculated readily by a standard
integration: 16
'Pm = -2

j:

12

[Pm(g) - Po]dg = -128a

64a 2
= ZK cRT exp(-ZKd)

2cRT j:12 exp(-2ZKg)dg


(6.17)

where 2g has been substituted for din Eq. 6.16 as a dummy variable of
integration.
Calculated values of Pm based on Eq. 6.9 and the integration of Eq. 6.12
agree semiquantitatively with experimentally determined reversible curves
of the swelling pressure of Na-montmorillonite versus the interplane
separation.j" These curves, which apply to the clay mineral suspended in
dilute solutions of NaCl, agree fairly well with the theoretical prediction
obtained with either O"p(= -0.118 C'm- 2) or 1/1(0) (= -0.25 V) held
fixed. The two theoretical curves are identical for d > 1 nm. A more direct
test of the DDL model of the electrostatic force is shown in Fig. 6.3. 21 The
graphs are semilogarithmic plots of the repulsive force per unit radius
(potential energy per unit area) between two muscovite cylinders in
contact with a solution of KN0 3 . The force between the two mica planes is
measured through the deflection of a spring attached to one of them, and
the interplane distance is measured by interferometry. The lines through
the data points in Fig. 6.3 represent Eq. 6.17 (with Z = 1). The excellent
agreement between experiment and theory is convincing evidence for the
applicability of DDL theory in 1: 1 electrolyte solutions at low concentrations. At higher concentrations of KN0 3 (above 10 mol, m'-3), however,

209

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

104 .___-.....---,.---.--..,.--"""T""-~-_,_-_..,.-_,--.___-~-_r_...,

MUSCOVITE IN KN0 3
20C
pH6
O.lmol m- 3
o I mol m- 3

10

10

20

30

40

50

60

70

80

90

100

110

120

d (nrn)
Figure 6.3. Measured values of the force F between two muscovite cylinders of
radius R as a function of their separation d in aqueous solutions of KN0 3 The ratio
F/2TrR equals the potential energy per unit area for the interaction between the two
mica surfaces. 21

and even at low concentrations (0.1 mol, m-3) in 2: 1 electrolyte solutions,


the DDL model predictions fail to describe the observed force curves.P
This failure can be traced to the inherent inadequacy of Eq. 5.1 for
phyllosilicates suspended in moderately concentrated 1 : 1 electrolyte solutions and in any asymmetric electrolyte solution, as discussed in Sec. 5.1
(see Fig. 5.1).
When considered over a time
interval much longer than 10- s, the distribution of electronic charge in a
nonpolar molecule is geometrically spherical. However, on a time scale
comparable to or shorter than 10- 16 s (approximately the period of an
ultraviolet light wave), the charge distribution of a nonpolar molecule
exhibits significant deviations from spherical symmetry, taking on an
evanescent dipolar (or higher multipole) character. These deviations
fluctuate rapidly enough to average to zero when observed over, say,
10- 14 s, but they persist long enough to induce distortions in the charge
distributions of neighboring molecules. If two nonpolar molecules are
brought close together, each induces in the other a fluctuating dipolar
character and the correlations between these induced dipole charge
distributions do not average to zero over a long time period, even though
the individual dipole distributions themselves average to zero. On a very
short time scale. a nonpolar molecule creates a dipolar electric field of
intensity E ... fJ..I R'\ where fJ.1 is the magnitude of the instantaneous
THE VAN DER WAALS DISPERSION FORCE.
16

THE SURFACE CHEMISTRY OF SOilS

210

dipole moment of the molecule and R is the distance between the center of
the molecule and the point at which E is evaluated.P This electric field in
turn induces a dipole moment J-L2 = aE = aJ-LdR 3 in another nonpolar
molecule separated from the first by the distance R, where a is the
polarizability of the second molecule. The correlation between the two
instantaneous dipole moments produces the potential energy
V = - J-LIJ-L2/R 3 according to classical electrostatics. Therefore, the average potential energy of interaction is V = - aJ-LilR 6 and the corresponding
force is F = -6aJ-LIIR 7 . This force is known as the van der Waals
dispersion force. For small values of intermolecular separation, this force
can be large, although it cannot compete with the much stronger covalent
interactions. At large intermolecular separations it is much smaller than
the coulomb force.
Suppose that a nonpolar molecule confronts the planar surface of a solid
comprising N of the same molecules per unit volume. The van der Waals
dispersion potential energy for the interaction between the single molecule
and the solid can be calculated with the equatiorr'"
(6.18)
Equation 6.18 is interpreted as follows. The single molecule is positioned a
distance z from the planar surface and a distance R = [(z + ()2 + p 2F/2
from a point within a ring-shaped element of the solid (Fig. 6.4). The
volume of this element of solid is dr = pd({Jdpd( in cylindrical polar
coordinates. The van der Waals potential energy between the single
molecule and the molecules in the volume element is
C
dV(z) = -R 6 N dr

(ex> (ex>

V(z)

-2TTNC

Jo Jo

-TTNC (ex>

pdpd(
[(z + ()2 + p2j3

d(

Jo (z + ()4

-TTNC

6z 3

Now suppose that the single molecule is embedded in a solid just like the
one it confronts. Within a thin slab of thickness dz that lies parallel to the
planar surface of the solid. there are Ndz molecules per unit area. and so

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

211

ATOM

SOLID
Figure 6.4. Cylindrical polar coordinates to define the spatial relationship between
an atom located a distance R away from a ring-shaped element of volume located a
distance ( inside a solid with a planar surface.

the potential energy per unit area between two opposing solids a distance d
apart is23
CPvw(d)

=N
-

d""

-7r:

V(z)dz

A
121Td 2

C L"" ::

(6.20)

where A == (1TN)2C is called the Hamaker constant. Equation 6.20 shows


that the van der Waals dispersion potential energy (per unit area) falls off
as the inverse square of the distance separating two opposing planar
surfaces. Because it is additive for all molecules in the two solids, this
attractive interaction decreases with distance of separation much more
slowly than the interaction between an isolated pair of molecules.
A tacit assumption in the derivation of Eq. 6.18 is that the response of
one molecule to the distorted charge distribution of the other is effectively
instantaneous. If the intermolecular separation is small enough, this
approximation does not affect the calculation of V(z), but at some distance
of separation its effect is significant. In broad terms, this distance is reached
when the time required for the ftuctuating dipolar field of one molecule to

212

THE SURFACE CHEMISTRY OF SOILS

propagate to the other molecule and be returned becomes comparable with


the period of the fluctuation itself. Under these conditions, the correlations
between the two induced dipole distributions weaken and the potential
energy of interaction eventually becomes proportional to the inverse
seventh power of the intermolecular distance instead of the inverse sixth
power.P This potential energy corresponds to a force depending on the
inverse eighth power of the distance, the retarded van der Waals dispersion
force. The transition between the nonretarded and retarded van der Waals
forces is not usually abrupt, but occurs gradually over distances on the
order of tens of nanometers. The retarded van der Waals force between
the two planar surfaces a distance d apart can be calculated following the
procedures in Eqs. 6.18 to 6.20, with the result

'PRvw(d)

= - 3d 3

(6.21)

where B is the retarded Hamaker constant.


The theoretical calculation of the Hamaker constants in Eqs. 6.20 and
6.21 for the case of two opposing planar solid surfaces with an aqueous
solution between them is a difficult area of current research. 22 ,24 An
approximate equation for A pertaining to this physical situation can be
derived under the assumptions that the solids are dielectrics, that only the
liquid water between them affects the van der Waals force, and that the
solids and the water absorb ultraviolet radiation at a single characteristic
frequency. The resulting equation is22
3(n~ - n;)2hWo

(6.22)

where n, (i = w or s) is a refractive index of water or solid, lU o is the


characteristic angular frequency of absorption, and h = 1.0546 X
10- 34 J . s is the Dirac constant. With nw = 1.33, ns = 1.6 (typical for
phyllosilicates), and Wo = 2 X 1016 rad- s -1 as a typical angular frequency,
Eq. 6.22 leads to the estimate A = 1.9 X 10- 20 J.
Theoretical estimates of B can be developed in a similar fashion.P with
the result B = 10- 28 J. m. Figure 6.5 shows measured values" of the van
der Waals potential per unit area for muscovite surfaces contacting
aqueous solutions of KN0 3 or Ca(N03h at concentrations between 10 and
103 mol m -3. The experimental method was the same as described in
connection with Fig. 6.3, but with conditions arranged to suppress the
electrostatic force. The line through the data points represents Eq. 6.20
multiplied by 27T on both sides, with A = 2.2 0.3 x 10- 20 J. The good
agreement for d < 7 nm between theory and experiment with respect to
the d dependence suggests that the van der Waals force does not depend
on the type of electrolyte present or, in the range examined, on its
concentration. Further support for the conclusion that A does not depend
on the type of electrolyte comes from the agreement between the

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

213

o ....-----.-------.----::-........--....-0

-50

( \J

IE

-100

"'?

--::l
-

"'C

-ft

-150
MUSCOVITE

I:::

C\J

KN0 3

Co(N0 3)2

-200

-250 L..-_ _-'--..L


o
5

....L.-

10

---I

15

d (nrn)
Figure 6.5. Measured values of the van der Waals potential energy per unit area,
'Pvw(d) , for opposing muscovite surfaces separated by a distance d in aqueous
solutions of KN0 3 and Ca(N03h at concentrations from 10 to lcP mol m ~3 and at
pH 6. 21

measured value and the theoretical estimate (Eq. 6.22) of this parameter,
since the latter is based on the properties of water alone in the aqueous
phase. For interplane separations larger than 7 nm, Fig. 6.5 suggests that
the retarded van der Waals force is important. The scatter in the
experimental data prevented a conclusive test of Eq. 6.21, however.
The electrostatic and van der Waals dispersion
forces retain the common attribute of depending on the nature of liquid
water in the aqueous solution phase only through the macroscopic dielectric constant. In the case of the electrostatic force as exemplified in
Eq. 6.16, the only dependence on the properties of liquid water comes
through the parameter 1<, which, as shown in connection with Eq. 5.11, is a
function of the bulk (zero-frequency) dielectric constant, D. Similarly, for
THE SOLVATION FORCE.

214

THE SURFACE CHEMISTRY OF SOilS

the nor-.retarded Hamaker constant represented by Eq. 6.22, the standard


equality between the square of the refractive index and the dielectric
constarst at optical frequencies'f can be applied to show that the van der
Waals dispersion force also depends only on the properties of liquid water
as a boll: dielectric continuum.
The @bsence of any parameters referring directly to the molecular nature
of liqui,d water is certainly a simplifying feature of the electrostatic and van
der w.-als dispersion forces, but it is evident that this simplicity ceases to
be realistic at some point as two planar phyllosilicate surfaces are brought
closer 2lnd closer together. For example, during quasicrystal formation by
unit layers of montmorillonite bearing exchangeable Ca2+ cations, one can
imagine that the competition between the repulsive electrostatic force and
the att ractive van der Waals force will, along with random thermal
motions. determine the behavior of two siloxane surfaces approaching one
another from a distance of separation larger than 10 nm. However, at the
interplelIl-ar separation distance of 0.95 nm, which is characteristic of the
outer-sf'bere surface complex in Fig. 6.1, it can be expected that the force
require d to bring the particle surfaces closer together must have a
compooent that reflects the effort necessary to desolvate the exchangeable
Ca 2 + ctltions. Indeed, at distances between 10 nm and 0.95 nm, the force
bringing the siloxane surfaces into close proximity must displace water
molecules from the second solvation shell of Ca2+ cations (Sec. 2.2). When
the two surfaces collapse into the quasicrystal configuration, almost all of
these outermost solvating water molecules will have been ejected from the
interlayer region (Fig. 2.4), and the force required to accomplish this task
must tD some extent depend on the structure of the cation solvation
complex at the molecular level. 25
Based on qualitative arguments only, it is not possible to derive an
explicit relationship between the additional force required to desolvate
exchangeable cations and the interplanar separation distance. Since most
of the metal cations encountered on the surfaces of natural phyllosilicates
tend to form octahedral primary solvation complexes in aqueous solutions
(Sees. :2.2 and 2.3), it is likely that the additional force increases markedly
as the ioterplanar separation decreases below 1 nm, the characteristic
diameter of octahedral solvation complexes. For separations larger than
1 nm, it is possible that the additional force drops rapidly to zero if the van
der Waals dispersion force is strong enough to overcome the electrostatic
force and displace water molecules from secondary solvation shells of the
cations in the aqueous solution between the opposing phyllosilicate
surfaces. The effect of the finite diameter (0.29 nm) of the water molecules
on the decay of the additional force with distance should be to superimpose
an oscillatory feature that exhibits relative maxima at the mean positions of
the molecules projected along a normal extending from one opposing
surface to the other. 25
Qualitative evidence for the existence of a solvation force between
siloxane surfaces has been adduced from experiments on the compression

215

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

of Na-montmorillonite particles aligned perpendicularly to their crystallographic c axes in a NaCI background solution.j" In a solution of
0.1 mol, m -3 NaCl, the reversible experimental compression curve (pressure applied versus interplanar separation) for Na-montmorillonite was
observed to always lie significantly above the theoretical compression
curve, based on Eqs. 6.9 and 6.11 and the derivative of 'Pvw(d) in Eq. 6.20,
when the interplanar separation decreased from 5.0 to 1.8 nm. This
positive deviation of the experimental compression curve became even
larger after the clay adsorbed n-dodecyl hexaoxyethylene glycol
monoether, a nonionic surfactant compound that reduces both the electrostatic and the van der Waals force. The conclusion drawn was that an
additional force must be significant at interplanar separations below 5 nm.
As with the electrostatic and van der Waals dispersion forces, a more
precise characterization of the solvation force has come from direct
experimental measurements of the net force per unit radius (energy per
unit area) between opposing muscovite surfaces.j? Figure 6.6 shows this

Figure 6.6. The net force per unit radius between muscovite surfaces in contract
with an aqueous solution of KBr at pH 6.2. 28 The solid line represents the
contribution of electrostatic and van der Waals forces.

MUSCOVITE
0.5 mol m- 3 KBr

a::
<,
LL.

ELECTROSTATIC ~
PLUS VAN DER WAALS
FORCES

102

10

20

30

d(nm)

40

50

216

THE SURFACE CHEMISTRY OF SOILS

net force as a function of interplanar separation for two mica surfaces in


contact with a solution of 0.5 mol, m -3 KBr at pH 6.2. 28 The solid line in
the figure represents the combination of the electrostatic and van der
Waals forces according to Eqs. 6.9, 6.11, and 6.20, with 1/1(0) = -0.1 V
taken as a fixed surface potential. For separations of the planes smaller
than about 6 nm, there is a marked positive deviation of the experimental
data from the theoretical curve. The difference between the data and the
curve can be fit by regression to the equatiorr"
'Psolv(d)

a
21T exp( -dj8)

(6.23)

where a and 8 are empirical constants. From this kind of curve fitting, it is
found that a = 0.03 to 0.05 N 'm- 1 and 8 = 0.3 to 1.0 nm. 27 ,29 The
solvation force decays approximately exponentially with a decay length,
8, near 1 nm, in agreement with the qualitative expectations outlined
above.
Several qualitative properties of the solvation force between the siloxane
surfaces of muscovite have been determined experimentallyr"
1. In the temperature range 20 to 65C, there is no change in either a or 8
in Eq. 6.23. This lack of temperature dependence ofthe solvation force
is consonant with the very small temperature dependence of the Gibbs
energy change for cation solvation in aqueous solutions.
2. The solvation force diminishes as the concentration of the aqueous
solution between the mica surfaces decreases. When the concentration
drops below 0.1 mol, m -3, the solvation force begins to disappear
altogether.
3. The solvation force is not detectable when the principal cation in the
aqueous solution between the mica surfaces is the proton.
4. In the presence of 1: 1 electrolyte solutions at pH 5 to 6, the solvation
force makes an appearance at a threshold concentration, whose value
depends on the metal cation in the electrolyte. These threshold
concentrations are 60 (Li), 10 (Na), 0.04 (K), and 1 (Cs) mol m -3 for
Group IA metals. Thus a in Eq. 6.23 is both cation- and concentrationdependent.
These experimental properties of the solvation force are consistent with
the hypothesis that exchangeable cations, either in outer-sphere surface
complexes or in a diffuse swarm, determine the nature of the solvation
force between siloxane surfaces. (Exchangeable cations in inner-sphere
surface complexes can also playa role if they are partially solvated.) If the
proton is the principal exchangeable cation (e.g., when the pH or the
electrolyte concentration is very low), desolvation seems to require no
more than the presence of the van der Waals force. For monovalent metal
cations, the threshold concentration at which the solvation force appears
increases as the selectivity of the siloxane surface for the cation relative to

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

217

potassium ions decreases (Fig. 4.4). This inverse correlation suggests that a
certain fraction of the K+ cations in the siloxane ditrigonal cavities of the
muscovite surface must be replaced by metal cations in outer-sphere
complexes and/ or the diffuse ion swarm in order for the solvation force to
be manifest. Since the proton is always a third exchangeable cation when
K+ ~ M+ (M = Li, Na, or Cs) exchange occurs, the appearance of the
solvation force should also depend on the selectivity of muscovite for the
proton in the ternary system K+-M+-H+-muscovite. Clearly, the chemical master variables that determine the appearance of the solvation force
are the composition of the aqueous solution between the mica surfaces and
the selectivity coefficients for K+-M+, K+-H+, and M+-H+ exchange.
Since the solvation force is defined as the difference between the observed
net force and the net van der Waals plus electrostatic force, as calculated
with Eqs. 6.9,6.11, and 6.20, it follows from the foregoing discussion that
Eq. 6.12 should be integrated in the context of an appropriate surface
complexation model (e.g., the triple layer model) to provide the most
accurate theoretical estimate of the electrostatic force. 30
6.3. THE STABILITY OF SOIL COLLOIDAL SUSPENSIONS

A suspension of soil colloids is said to be stable if the average particle size


of the colloids does not increase at a significant rate to the point when
gravitational settling can occur.P If a soil colloidal suspension is unstable,
it is said to undergo coagulation as its constituent particles interact to form
larger particles. When coagulation produces a relatively loose, open
network of linked particles in a structure that is often ephemeral, the
process is called flocculation. When more permanent, compact structures
are produced instead, the process is called aggregation. The smallest
concentration of electrolyte, in moles per cubic meter, at which a soil
colloidal suspension begins to undergo rapid coagulation is called the
critical coagulation concentration (ccc). The value of the ccc, in general,
depends on both the nature of the colloidal particles and the composition
of the aqueous solution in which they are suspended. The measurement of
the ccc entails the preparation of dilute (less than 3 kg' m- 3 solids
concentration) suspensions in a series of solutions of increasing electrolyte
concentration. After 1 hour of shaking and a standing period of 24 hours,
the coagulated suspensions show a clear boundary separating the settled
solid mass from an aqueous solution phase, and the ccc can be bracketed
between two values determined by the largest electrolyte concentration at
which coagulation does not occur and the smallest at which it does. 32
Colloidal stability plays an important role in the maintenance of the soil
aggregate structures on which the penetrability of the soil surface and the
permeability of the soil profile depend." The study of soil colloidal
stability has not yet produced exact, quantitative theories, but there have
begun to emerge general relationships between stability, interparticle
forces, and surface chemistry that are of predictive value.

218

THE SURFACE CHEMISTRY OF SOILS

The empirical relationship concerning the ccc


first suggested by Schulze 100 years ago and generalized by Hardy in 1900
can be statedr'"
THE SCHULZE-HARDY RULE.

The critical coagulation concentration for a colloid suspended in an aqueous


electrolyte solution is determined by the ions with a charge opposite in sign to
that on the colloid and is proportional to an inverse power of the valence of the
ions.

The Schulze-Hardy rule is illustrated in Table 6.2 for five important soil
minerals." Mean values of the ccc extracted from published studies,
together with standard deviations reflecting the range of ccc values
reported, are presented in the table for ions whose absolute valence is
equal to 1 or 2. For the hydrous oxides, the particle charge is positive and
the coagulating ions are anions; for the phyllosilicates, the particle charge
is negative and the coagulating ions are cations. The ratio of ccc values for
IZI = 2 and !ZI = 1 is given in the fourth column of the table. These ratios
may be compared with a theoretical value of r 6 = 1/64 = 0.0156, which
is derived below. It is evident from Table 6.2 that the Schulze-Hardy rule
provides a semiquantitative prediction of relative ccc values for soil
colloidal suspensions. The inverse sixth power of the coagulating ion
valence appears to be a reasonable factor for relating ccc values.
A theoretical derivation of the Schulze-Hardy rule can be developed on
the basis of the interparticle forces described in Sec. 6.2. Each of the three
forces is associated with a potential energy that contributes additively to
the total potential energy between two planar particle surfaces a distance d
apart. If <p(d) is the total potential energy per unit area of planar surface,
then
(6.24)
and if a symmetric electrolyte is present such that ZKd/2 remains large,
64a 2
A
<p(d) = ZK cRT exp( -ZKd) - 127Td2

+ 27T exp( -d/ 13) (6.25)

according to Eqs. 6.17,6.20, and 6.23. It is assumed in Eq. 6.25 that dis
Table 6.2. Critical coagulation concentrations for colloidal suspensions of soil
minerals35
Soil
mineral
Al hydrous oxide
Fe hydrous oxide
Illitic mica
Kaolinite
Montmorillonite

ccc (121 = 1),


mol-rn ?
50
11
48
10
8

9
2
11
4
6

ccc (121 = 2),


mol-m ?
0.5 0.2
0.21 0.01
0.14 0.02
0.3 0.2
0.12 0.02
Simple DLVO theory:

ccc(121 = 2)
ccc (121 = 1)
0.010
0.019
0.003
0.030
0.015
0.0156

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

219

large enough to make Eq. 6.17 applicable but not so large as to require a
retarded van der Waals dispersion force (Eq.6.21). Since Eq.6.17
approximates 'Pm(d) within about 20 per cent for d > I/ZK and 'Pvw(d)
begins to take on retarded character for d above 7 nm, 17 ,21 the range of
d-values over which Eq. 6.25 should be applicable is (at 298 K) 9.6/ zJC
< d < 7 nm, where c is the electrolyte concentration in moles per cubic
meter. If the ccc values in the second column of Table 6.2 are substituted
for c, the lower limit of d lies between 1 and 4 nm. If the ccc values in the
third column of the table are used, the required lower limit of d generally
exceeds the upper limit of 7 nm (e.g., for c = 0.5 mol m -3, the lower limit
of d is 6.8 nm). Therefore, 'Pvw(d) is better represented by Eq. 6.21 than
by Eq. 6.20 in this case. On the other hand, 'Pm(d) calculated from the
Poisson-Boltzmann equation is not very accurate for bivalent ions, and the
value of the retarded Hamaker constant in Eq. 6.21 is not known
accurately for soil minerals. These facts suggest that Eq. 6.25 can serve as
well as any available expression to illustrate the relationship between the
ccc and ionic valence in a simple and qualitative fashionr'"
Since 'P(d) comprises both positive and negative terms, it is expected to
show a relative maximum at some d value for many possible choices of
Z, K, a, A, a, and 8. This kind of mathematical behavior is illustrated in
Fig. 6.7 for the case Z = 1, K = 0.329 nm", a = 0.462, A = 2.210- 20 J,
a = 0.05 J . m -2, and 8 = 1 nm. These values are appropriate for a
stable suspension of illitic mica in 10 mol-rn"? NaCI at 298 K. 29 Figure 6.7
Figure 6.7. Total potential energy per unit area for the interaction of two charged
parallel phyllosilicate surfaces, with 10 mol m -3 NaCl interposed between them,
according to Eq. 6.25. See also the solid curve in Fig. 6.6.

Interplanar separation,

-4

THE SURfACE CHEMISTRY OF SOilS

220

shows the characteristic behavior of ({J(d) , wherein a pronounced minimum


occurs for very small values of d followed by the maximum and at least one
more relative extremum. (See also the solid curve in Fig. 6.6.) If two more
extrema occur, the one at a finite d value is a shallow minimum termed the
secondary minimum. (The primary minimum is the very deep one at small
d values.) Sometimes flocculation is associated with the secondary minimum and aggregation with the primary minimum.V
On the hypothesis developed by Derjaguin, Landau, Verwey, and
Overbeek (DLVO),38 a colloidal suspension becomes rapidly unstable if
the maximum value of ({J(d) is small relative to the random thermal energy
of the colloidal particles. This hypothesis forms the basis of the DLVO
theory of colloidal stability. With the approximate expression for ({J(d) in
Eq. 6.25, one can apply the DLVO hypothesis in the form

({J(d)

(a({Jjad)ccc

=0

(c = ccc)

(6.26)

to derive a relationship between the ccc and Z. Equations 6.26 state that
both ((J(d) and its first derivative with respect to d vanish when c in
Eq. 6.25 equals the ccc. The first condition implies the equation

64a2
A
ZK ccc RT exp( -ZKcd) = 127Td 2
c

a
-

127T exp( -dI5)

(6.27a)

whereas the second condition on ((J(d) yields the expression


A

64 a2 ccc RT exp( -ZKcd) = 67Td3

al5
27T exp( -dl({J)

(6.27b)

where Kc = J {3 ccc = 0.1041 Jccc nm " at 298 K with the ccc in moles per
cubic meter. The right side of Eq. 6.27a must be positive in order that the
value of the ccc remain positive. This requirement is met for sufficiently
small values of d (since a? grows arbitrarily large and exp( -dl 5) remains
finite as d goes to zero) as well as for sufficiently large values of d (since d- 2
approaches zero more slowly than exp( -dl 5) as d goes to infinity. In some
intermediate range of d values, the right side of Eq. 6.27a can be negative
if the hydration force dominates the van der Waals dispersion force. The
colloidal particles maintain a stable suspension so long as this condition
persists.
In DLVO theory, it is customary to assume that the electolyte concentration and suspended particle configuration are such that the van der
Waals dispersion force dominates the hydration force to the extent that the
latter can be neglected.l" Under these conditions of electrolyte concentration and particle configuration (which differ for different kinds of soil
colloid and background electrolyte), the second term on the right side of
Eq. 6.27 is dropped and the two expressions are divided to derive the result

z-, =

(6.28)

The substitution of Eq, 6.2H into Eq. 6.27u (without the term in u) thcn

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

221

gives the equation

.3/ _ (30721T
a RT) -3
KC eee e2
A Z
2

or

2RT)2
_(30721Ta
eee Z
2

e f3

3/2

-6

(6.29)

Equation 6.29 represents the Schulze-Hardy rule according to a simplified


application of the DLVO theory. It follows from this equation that
eee(Z = 2)/eee(Z = 1) = 2- 6 = 0.0156, in good agreement with the
trend in the fourth column of Table 6.2. However, this good agreement
must be considered in large measure the result of a fortuitous cancellation
of the effects of the physical factors neglected when Eq. 6.25 is invoked
without the term in fPsolv(d). Although the interplay of electrostatic and
van der Waals dispersion forces as represented in Eqs. 6.17 and 6.20 is
sufficient to derive a proportionality between the ccc and a power of the
ionic valence via the DLVO hypothesis in Eq. 6.26, the detailed numerical
relationship found in Eq. 6.29 is not accurate (it predicts impossibly large
ccc values) and is not improved by a better representation of fPm(d) based
on the exact solution of the Poisson-Boltzmann equation. 17 The derivation
of an equation to predict the ccc values in Table 6.2 awaits greater
precision in the determination of both the parameters and the distance
dependence in the component potentials of fP(d).
According to the DLVO theory, the rapid
coagulation of a colloidal suspension is induced by a reduction in both the
magnitude and the range of the repulsive electrostatic force as the
concentration of electrolyte increases. The essential physical soundness of
this conceptual view has been demonstrated experimentally in studies of
the behavior of montmorillonite suspended in either Liel or NaC!. The
colloidal particles in this case are individual unit layers (Table 6.1), and the
principal electrostatic force between them emanates from their siloxane
surfaces. When the electrolyte concentration is very low (less than
0.01 mol-rn"), the decay length ofthe electrostaticforce, K- 1 in Eq. 6.17,
is very large (greater than 60 nm), even in suspensions whose clay
concentration is around 40 kg-rn", Under the conditions, both the van
der Waals dispersion force and the hydration force are negligible and the
electrostatic force is powerful enough to order the montmorillonite particles into parallel stacking along a direction perpendicular to their crystallographic e axis. This arrangement of the clay particles, known as a tactoid,
has been found through small-angle neutron scattering experiments on
suspensions of Li-montmorillonite.r" The interplanar separation distances
varied from about 40 nm at a clay concentration of 68 kg m -3 to about
120 nm at a clay concentration of 20 kg- m -3. These large interplanar
separations and the repulsive force that produces them distinguish the
tactoid clearly from the quaslcrystal, which is produced by attractive forces
COMPLEXATION REACTIONS.

I'

..

222

THE SURFACE CHEMISTRY OF SOILS

The characteristic behavior of the electrostatic force in bringing about


parallel arrangements of montmorillonite particles has also been verified at
electrolyte concentrations approaching and exceeding the ccc of Namontmorillonite suspended in NaCl. 40 However, a different situation exists
when the coagulating ions have a valence different from unity or, in
general, when these ions can engage significantly in complexation reactions, either with surface functional groups or with other ions in the
aqueous solution phases. Complexation reactions in the aqueous solution
phase tend to produce what is termed the antagonistic effect;" wherein the
ccc increases as the concentration of an added complexing ion increases.
For example, if Ca2+ cations alone produce the coagulation of a negatively
charged colloid according to the DLVO conceptualization, then the ccc,
defined conventionally in terms of the total concentration of the metal."
must increase as the concentration of an added anion that can form soluble
complexes with the metal increases. In other words, the presence of a
complexing anion, say SO~-, reduces the concentration of the free ionic
species, Ca2+, through the formation of the complex CaSO~, and a larger
total calcium concentration is required to bring the concentration of the
species Ca 2+ up to the level required to produce coagulation." This effect,
of course, occurs equally well with complexing cations added to a
suspension of positively charged colloid and coagulating anions.
The formation of soluble complexes by coagulating ions reduces the
effectiveness of these ions in diminishing the electrostatic force, but the
fundamental mechanism that brings about coagulation is not changed. On
the other hand, if the coagulating ions can form surface complexes with the
suspended colloidal particles, the nature of the coagulation process does
change and the DLVO hypothesis becomes inappropriate. The formation
of surface complexes alters the total particle charge density, up, directly,
whereas the formation of the compact diffuse double layer inherent to the
DLVO hypothesis only screens the total particle charge density to reduce
its long-range effect. If the surface functional group is the siloxane
ditrigonal cavity, inner-sphere complex formation with monovalent metal
cations can reduce the total particle charge sharply. Since this type of
surface complex formation is more likely, it follows that, as the soft Lewis
acid character of the metal cation increases (Sec. 4.3), Cs+ should be more
effective than Na+, for example, at reducing up and thereby promoting
coagulation. Experimental evidence supporting this hypothesis comes from
the observation that the ccc of a dilute Na-montmorillonite suspension
(0.25 kg' m -3) at pH 6 is 2.1 mol, m -3 in NaN0 3 , whereas the ccc of a
Cs-montmorillonite suspension at the same clay concentration and pH is
0.79 mol m -3 in CsN0 3 42 The lower cccfor Cs-montmorillonite evidently
reflects a larger number of inner-sphere surface complexes formed per unit
area and, therefore, a lower up'
In the case of bivalent metal cations, siloxane ditrigonal cavities tend to
form outer-sphere surface complexes leading to quasicrystals, as described
in Sec. 6.1. Since two ditrigonal cavities are involved, quasicrystal forma-

SURFACE CHEMICAL ASPECTS OF SOil COllOIDAL STABILITY

223

tion also produces an electrically neutral surface, but it is the internal


surface of a quasicrystal that has its total charge density reduced. Experiments on the coagulation of Ca-montmorillonite suggest that outer-sphere
complexation of Ca2+ is an aggregation process that follows a flocculation
process induced by Ca2+ cations participating in the surface chargescreening mechanism hypothesized in the DLVO theory.P In this case,
outer-sphere complexation is a gradual particle rearrangement phenomenon that may require several days' time. When initiated from a stable
suspension, however, outer-sphere complexation and quasicrystal formation can be very rapid, as pointed out in Sec. 6.1.
When the principal surface functional group is the inorganic or organic
hydroxyl group, colloid stability can be affected strongly by the pH value,
since inner-sphere complexes with protons and hydroxide ions are formed.
A soil colloidal suspension containing particles bearing surface hydroxyl
groups (e.g., hydrous oxides or kaolinite) tends to coagulate at the PZC
regardless of the background electrolyte concentration. In the absence of
surface complexes involving ions other than H+ or OH-, the PZC
coincides with the PZNPC and coagulation depends on the balance of
surface charge between protonated and dissociated functional groups as
expressed through O"H; In the presence of other complex-forming ions, O"p
is the determining property for rapid coagulation and the effects of the
ions in the aqueous solution phase must be evaluated as described in
Sees. 3.1 and 3.2. As a general rule, the existence of ions that can form
surface complexes significantly can be detected by examining the ccc as a
function of the colloid concentration in the suspension. If the DLVO
mechanism is the principal cause of coagulation, the ccc is essentially
independent of the colloid concentration-at least over a severalfold
change-whereas if surface complexation is the principal cause, the ccc
tends to increase with the colloid concentration, since the surface complexation capacity is also increased.t" If the complexed ion is multivalent,
surface complexation can result in a reversal of the sign of O"p, as described
in Sec. 4.3 for metal cations. When this happens, the ions in the aqueous
solution phase that previously were of the same charge sign as the colloidal
particles become potential coagulating ions. The mechanism of any
subsequent coagulation induced by these ions can be either surface
charge-screening or surface complexation.
When polymer ions form surface complexes with soil colloidal particles,
stability depends on stereochemistry as well as on surface charge density. 45
If the extent of polymer adsorption is small, a soil colloidal suspension may
be coagulated at a lower concentration of an added noncomplexing
electrolyte than in the absence of the polymer. In this situation, the
addition of electrolyte brings the colloidal particles closer and closer
together until the polymer chains can form bridges among them, inducing
coagulation. Since the polymer bridging can occur with the particles farther apart than the separation required to make the particles' own
van der Waals forces effective. coagulation can occur at lower electrolyte

224

THE SURFACE CHEMISTRY OF SOILS

Table 6.3. Factors affecting the stability of soil colloidal suspensions


Factor
Electrolyte
concentration
pH value
Surface complexes
with small ions
Surface complexes
with polymer
ions

Effects

Promotes
stability

Promotes
coagulation

Extent of diffuse
double layer;
solvation force
Changes UH
Changes up

When increased

When decreased

pH = PZC
up = 0

pH

Changes up and/or
particle association

With polymer
bridges

By electrostatic
repulsion

up

=f PZC
=f 0

concentrations. If polymer adsorption is significant, coagulation by polymer bridging may take place at extremely low electrolyte concentrations
or the colloidal suspension may be stabilized by the electrostatic force
between the coatings of adsorbed polymers. Which phenomenon occurs
depends on pH, on electrolyte concentration, and on the configuration of
the adsorbed polymer ion.
The principal surface chemical factors that determine the stability of soil
colloidal suspensions are summarized in Table 6.3. Surface reactions
affect colloid stability through changes in the strength of the repulsive
electrostatic and solvation forces and, if macromolecules are involved,
through changes in particle association mechanisms. Coagulation is the
result of a reduction in the efficacy of the repulsive electrostatic and
solvation forces, whether through charge-screening, surface complexation, or stereochemically induced particle bridging.
NOTES
1. The concept of the quasicrystal is distinguished from that of the tactoid in
J. P. Quirk and L.A.G. Aylmore, Domains and quasicrystalline regions in clay
systems, Soil Sci. Soc. Arn. I. 35: 652 (1971).
2. K. Norrish and J. P. Quirk, Crystalline swelling of montmorillonite, Nature
173: 225 (1954). A. M. Posner and J. P. Quirk, The adsorption of water from
concentrated electrolyte solutions by montmorillonite and illite, Proc. Royal
Soc. 278A:35 (1964). A. M. Posner and J. P Quirk, Changes in basal spacing
of montmorillonite in electrolyte solutions, I. Colloid Sci. 19: 798 (1964).
3. See, e.g., G. Sposito and R. Prost, Structure of water adsorbed on smectites,
Chern. Rev. 82: 553 (1982).
4. D. J. Cebula, R. K. Thomas, and J. W. White, Small-angle neutron scattering
from dilute aqueous dispersions of clay, l.eS. Faraday 176: 314 (1980).
R. Hight, W. L. Higdon, and P. W. Schmidt, Small-angle scattering study of
sodium montmorillonite clay suspensions, J. Chern. Phys. 33: 1656 (1960).
R. Hight, W. T. Higdon, H.C.H. Darley, and P. W. Schmidt, Small-angle
X-ray scattering from montmorillonite clay suspensions: II, J. Chern. Phys. 37:
502 (1962).

SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY

225

5. I. Shomer and U. Mingelgrin, A direct procedure for determining the number


of plates in tactoids of smectites: The Na/Ca montmorillonite case, Clays and
Clay Minerals 26: 135 (1978).
6. I. Shainberg and H. Otoh, Size and shape of montmorillonite particles saturated with Na/Ca ions (inferred from viscosity and optical measurements),
Israel J. Chem. 6:251 (1968). L. L. Schramm and J.C.T. Kwak, Influence of
exchangeable cation composition on the size and shape of montmorillonite
particles in dilute suspension, Clays and Clay Minerals 30: 40 (1982).
7. A. Banin and N. Lahav, Particle size and optical properties of montmorillonite in suspension, Israel J. Chem. 6:235 (1968). L. L. Schramm and J.C.T.
Kwak, op. cit." J. E. Dufey and A. Banin, Particle shape and size of two
sodium calcium montmorillonite clays, Soil Sci. Soc. Am. J. 43: 782 (1979).
8. D. G. Edwards, A. M. Posner, and J. P. Quirk, Repulsion of chloride ions by
negatively charged clay surfaces. Part 2. Monovalent cation montmorillonites,
Trans. Faraday Soc. 61: 2816 (1966). Part 3. Di- and trivalent cation clays,
Trans. Faraday Soc. 61:2820 (1966). L. L. Schramm and J.C.T. Kwak,
Interactions in clay suspensions: The distribution of ions in suspension and the
influence of tactoid formation, Colloids Surfaces 3: 43 (1982).
9. D. G. Edwards et al., op. cit.,s Part 3, Fig. 1.
10. Light transmission: L. L. Schramm and J.C.T. Kwak, op. cit:" Electrophoretic
mobility: P. Bar-On, I. Shainberg, and I. Michaeli, Electrophoretic mobility of
montmorillonite particles saturated with Na/Ca ions, J. Colloid Interface Sci.
33:471 (1970). Intrinsic viscosity: I. Shainberg and H. Otoh,op. cit.6 Chloride
exclusion volume: J. E. Dufey, A. Banin, H. G. Laudelout, and Y. Chen,
Particle shape and sodium self-diffusion coefficient in mixed sodium-calcium
montmorillonite, Soil Sci. Soc. Am. J. 40: 310 (1976).
11. I. Shainberg and A. Kaiserman, Kinetics of the formation and breakdown of
Ca-montmorillonite tactoids, Soil Sci. Soc. Am. J. 33: 547 (1969).
12. I. Shainberg and W. D. Kemper, Electrostatic forces between clay and cations
as calculated and inferred from electrical conductivity, Clays and Clay Minerals
41: 117 (1966). I. Shainberg, J. D. Oster, and J. D. Wood, Electrical conductivity of Na/Ca-montmorillonite gels, Clays and Clay Minerals 30: 55 (1982).
13. G. A. O'Connor and W. D. Kemper, Quasicrystals in Na-Ca systems, Soil Sci.
Soc. Am. J. 33:464 (1969).
14. R. Keren and I. Shainberg, Water vapor isotherms and heat of immersion of
Na/Ca-montmorillonite systems. II: Mixed systems, Clays and Clay Minerals
27: 145 (1979).
15. L. L. Schramm and J.C.T. Kwak, Thermochemistry of ion exchange and
particle interaction in clay suspensions, Can. J. Chem. 60: 486 (1982).
16. That ideal solution behavior is essential to the derivation of Eq. 6.7a has been
shown in W. Olivares and D. A. McQuarrie, Interaction between electrical
double layers, J. Phys. Chem. 84: 863 (1980).
17. See Chapter IV in E.J.W. Verwey and J.T.G. Overbeek, Theory of the
Stability of Lyophobic Colloids. Elsevier, Amsterdem, 1948. E. P. Honig and
P. M. Mul, Tables and equations of the diffuse double layer repulsion at
constant potential and at constant charge, J. Colloid Interface Sci. 36: 258
(1971).
18. D. Y.C. Chan, R. M. Pashley, and L. R. White, A simple algorithm for the
calculation of the electrostatic repulsion between identical charged surfaces in
electrolyte, J. Collotd lnterfac Sci. 77:2H3 (1980).

226

THE SURFACE CHEMISTRY OF SOilS

19. T. W. Healy, D. Chan., and L. R. White, Colloidal behaviour of materials


with ionizable group surfaces, Pure Applied Chem. 52: 1207 (1980).
20. L. M. Barclay and R. H. Ottewill, Measurement of forces between colloidal
particles, Spec. Disc. Faraday Soc. 1: 138 (1970). See also B. P. Warkentin,
G. H. Bolt, and R. D. Miller, Swelling pressure of montmorillonite, Soil Sci.
Am. J. 21:495 (1957). I. Shainberg, E. Bresler, and Y. Klausner, Studies on
Na/Ca montmorillonite systems. 1. The swelling pressure, Soil Sci. 111:214
(1971).
21. J. N. Israelachvili and G. E. Adams, Measurement of forces between two
mica surfaces in aqueous electrolyte solutions in the range 0-100 nm, J.C.S.
Faraday Trans. 174: 975 (1978).
22. J. N. Israelachvili and D. Tabor, Van der Waals forces: Theory and experiment, Prog. Surface Membrane Sci. 7: 1 (1973. See also J. N. Israelachvili and
D. Tabor, The measurement of van der Waals dispersion forces in the range
1.5 to 130 nm, Proc. Royal Soc. (London) 331A:19 (1972).
23. This kind of standard calculation is described in Chap. 10 of P. C. Hiemenz,
Principles of Colloid and Surface Chemistry. Marcel Dekker, New York, 1977.
24. J. N. Israelachvili, The calculation of van der Waals dispersion forces between
macroscopic bodies, Proc. Royal Soc. (London) 331A:39 (1972). See also
Chap. 7 in J. Mahanty and B. W. Ninham, Dispersion Forces. Academic
Press, London, 1976.
25. Similar concepts of the solvation force are presented in B. W. Ninham,
Long-range vs. short-range forces. The present state of play. J. Phys. Chem.
84: 1423 (1980).
26. L. M. Barclay and R. H. Ottewill, op. cit.2o
27. J. N. Israelachvili and G. E. Adams, op. cit.21
28. R. M. Pashley, Hydration forces between mica surfaces in aqueous electrolyte
solutions, J. Colloid Interface Sci. 80: 153 (1981).
29. R. M. Pashley, DLVO and hydration forces between mica surfaces in Li+,
Na+, K+, and Cs+ electrolyte solutions: A correlation of double-layer and
hydration forces with surface cation exchange properties, J. Colloid Interface
Sci. 83: 531 (1981).
30. A rudimentary form of the triple layer model has been applied in this way by
R. M. Pashley, op. cit.29
31. See, e.g., D. H. Everett, Manual of Symbols and Terminology for Physicochemical Quantities and Units. Appendix II: Definitions, Terminology and Symbols in Colloid and Surface Chemistry. Butterworths, London, 1972.
32. See, e.g., p. 23 in H. van Olphen, An Introduction to Clay Colloid Chemistry,
2nd edn. Wiley, New York, 1977.
33. J. P. Quirk, Some physico-chemical aspects of soil structural stability-A
review, in Modification of Soil Structure (W. W. Emerson, R. D. Bond, and
A. R. Dexter, eds.) Wiley, Chichester, U.K., 1978. Y. Chen and A. Banin,
Scanning electron microscope (SEM) observations of soil structure changes
induced by sodium-calcium exchange in relation to hydraulic conductivity, Soil
Sci. 120:428 (1975). Y. Chen, J. Tarchitzky, J. Brouwer, J. Morin, and
A. Banin, Scanning electron microscope observations on soil crusts and their
formation, Soil Sci. 130:49 (1980).
34. J.T.G. Overbeek, The rule of Schulze and Hardy, Pure Appl. Chem. 52: 1151
(1980).
35. E.J.W. Verwey and J.T.G. Overbeek, op. cit. ,Itt p. 9. S. L. Swartzen-Allen

SURfACE CHEMICAL ASPECTS Of SOIL COLLOIDAL STABILITY

36.
37.

38.

39.
40.

41.

42.
43.

44.
45.

227

and E. Matijevic, dolloid and surface properties of clay suspensions. III.


Stability of montmorillonite and kaolinite, J. Colloid Interface Sci. 56: 159
(1976) J. D. Oster, I. Shainberg, and J. D. Wood, Flocculation value and gel
structure of sodium/calcium montmorillonite and illite suspensions, Soil Sci.
Soc. Am. J. 44:955 (1980), and the references cited in these papers.
Somewhat more refined derivations of this relationship are presented in E. P.
Honig and P. M. Mul, op. citY and in J.T.G. Overbeek, op. cit.34
R.S.B. Greene, A. M. Posner, and J. P. Quirk, A study of the coagulation of
montmorillonite and illite suspensions by calcium chloride using the electron
microscope, in W. W. Emerson et aI., op. citY
B. V. Derjaguin and L. Landau, A theory of the stability of strongly charged
lyophobic sols and the coalescence of strongly charged particles in electrolytic
solutions, Acta Phys.-Chim. USSR 14:633 (1941). E.J.W. Verwey and J.T.G.
Overbeek, op cit.16
D. J. Cebula and R. H. Ottewill, Neutron diffraction studies on lithium
montmorillonite-water dispersions, Clays and Clay Minerals 29:73 (1981).
I. C. Callaghan and R. H. Ottewill, Interparticle forces in montmorillonite
gels, Faraday Disc. Chem. Soc. 57: 110 (1974). E. Frey and G. Lagaly,
Selective coagulation and mixed layer formation from sodium smectite solutions, Proc. Int. Clay Conf. 1978, p.l31 (1979). B. Rand, E. Pekenc, J. W.
Goodwin, and R. W. Smith, Investigation into the existence of edge-face
coagulated structures in Na-montmorillonite suspensions, J.C.S. Faraday 176:
225 (1980).
The effect of complexation reactions on colloid stability is described, with
many examples, in E. Matijevic, Colloid stability and complex chemistry, J.
Colloid Interface Sci., 43:217 (1973). See also E. Matijevic, The role of
chemical complexing in the formation and stability of colloidal dispersions, J.
Colloid Interface Sci. 58: 374 (1977).
See Figs. 2 and 6 in S. L. Swartzen-Allen and E. Matijevic, op cit.35
R.S.B. Greene, A. M. Posner and J. P. Quirk, Factors affecting the formation
of quasicrystals of montmorillonite, Soil Sci. Soc. Am. J. 37: 457 (1973). R.S.B.
. 37
1 op. cit.
G reene et a.,
W. Stumm, C. P. Huang, and S. R. Jenkins, Specific chemical interaction
affecting the stability of dispersed systems, Croatica Chem. Acta 42: 223 (1970).
For an introductory review, see Sec. 2.4 in B.K.G. Theng, Formation and
Properties of Clay-Polymer Complexes. Elsevier, Amsterdam, 1979.

FOR FURTHER READING


W. W. Emerson, R. D. Bond, and A. R. Dexter, Modification of Soil Structure.
Wiley, Chichester, U.K., 1978. The first five chapters of this symposium publication provide details on many aspects of the surface chemical features of soil
colloidal stability.
J. N. Israelachvili, Forces between surfaces in liquids, Advan. Colloid Interface
Sci. 16:31 (1982). An advanced-level summary covering the topics discussed in Sec.
6.2 of the present book.
B. W. Ninham, Long-range vs, short-range forces: The present state of play, J.
Phys. Chem, 84: 1423 (1980). This lively account of the current state of understanding of interparticle forcefl is mUflt reading to follow Sec. 6.2.

228

THE SURFACE CHEMISTRY OF SOILS

B. W. Ninham, Hierarchies of forces: The last 150 years, Advan. Colloid. Interface
Sci. 16: 3 (1982). Another must-reading article, this one dealing with the limitations
of interparticle force theories that assume the liquid phase to be a continuum
dielectric.
H. van Olphen, An Introduction to Clay Colloid Chemistry, 2nd edn. Wiley, New
York, 1977. Chapters 2, 3, 4, and 7 of this standard monograph form a useful
adjunct to the present chapter as regards phyllosilicate suspensions.
J.T.G. Overbeek, Strong and weak points in the interpretation of colloid stability,
Advan. Colloid Interface Sci. 16: 17 (1982). An overview of the status of DLVO
theory by one of the Starting Four.
E.J.W. Verwey and J.T.G. Overbeek, Theory of the Stability of Lyophobic
Colloids. Elsevier, Amsterdam, 1948. Part II of this classic monograph should be
read by the mathematically minded as a comparison to Sec. 6.2.

SELECTED PHYSICAL CONSTANTS*

Avogadro constant

NA

6.02252 x 1023 mol"!

Boltzmann constant

kB

1.38054 x 10- 23 J K- 1

Diffuse double layer constant

f3

1.084 x 1016 m mol"!

Faraday constant

9.64870 x 104 C mol "

Molar gas constant

8.3143 J K- 1 mol"!

Permittivity of vacuum

8.85419 x 10- 12 C2

r '

m- 1

D. D. Wagman, W. H. Evans, V, B, Parker, R. H. Schumm, I. Halow, S. M. Bailey, K. L.


Churney, and R. L. NUllall, The NBS tables of chemical thermodynamic properties, J. Phy.
Chem. Rr!. DUlu II,Supp. 2:1 (19H2).

INDEX

Adsorbate, 26
Adsorbed water, 58, 69
excess acidity, 71
phyllosilicates, 70
relation to ionic potential, 28
Adsorption, 25,29, 113, 122
defined, 113
isotherm, 116-117,122
kinetics, 127-28, 140
metal cations, 128
negative, 31,106,109
organic matter, 143
oxyanions, 138
pH effect, 134
precipitation versus, 122-28
surface excess, 113
Adsorption edge, 135, 139, 175, 183
Adsorption envelope, 139-40
Adsorption isotherm, 116-17
Adsorptive, 26
Aggregation, 217
Aluminol group, 18, 40
Anion exchange, 143-44
Anion exchange capacity, 36, 80

Babcock model, 109


Balance of surface charge, 178-81
general equation, 79
model equation, 179,189-90,192
BET equation, 27-28

Cation bridalna, 143,145-46


Cation exchanae capacity, 36, 80
Ccurve taotherm, 117

Chlorite, 6-7
weathering sequence, 21
Coagulation, 217
Colloidal stability, 217-24
and complex formation, 221-24
DLVOtheory, 220
factors affecting, 224
Constant capacitance model, 169-77
anion adsorption, 176-77
metal adsorption, 174-76
protonation, 170-74
PZNPC, 173
Coprecipitation, 8, 125
CPB method, 29
Critical coagulation concentration, 217,
222
complexation effects, 222
DLVO theory, 220-21
measurements, 218
polymer effects, 223
Schulze-Hardy rule, 218

Diffuse double layer (DDL) theory,


154-62
analytical solutions, 158
electrostatic force, 206
Monte Carlo simulations, 160-62
triple layer model, 178
Diffuse layer charge, 79,81,98,100,159,
178,207
Distribution coefficient, 27, 118-19
DLVO theory, 220
Donnan potential, 91
D.truclure, 48,54,59,67

232
Electrochemical potential, 88-90, 186
Electrokinetic phenomena, 94-106
convection current, 95
general equation, 97
Electrokinetic plane of shear, 94, 105
Electro-osmosis, 101, 103
Electrophoretic mobility, 83,97,99-100
defined, 97
quasicrystal formation, 203
relation to PZC, 98
surface complexes, 99-100,175-76
surface hydrolysis, 136-37, 183-84
Electrostatic force, 205-9
DDL model, 206-9
measurements, 208-9
Elovich equation, 127-28,140-41
Exclusion volume, 31,107,202
calcium montmorillonite, 202
defined, 31, 107
diffuse double layer theory, 31-33
molecular interpretation, 107
quasicrystal formation, 34-35,202-3
relation to specific surface area, 32, 107
sodium montmorillonite, 33-34

Flocculation, 217

Galvani potential, 91-92


Gibbs factor, 163-65
Gibbsite, 4-5
PZC, 84
surface charge, 40
surface hydroxyls, 5, 17
Goethite, 4-5
adsorption envelope, 140
ligand exchange, 140
surface area, 24
surface charge, 40
surface complex, 16-17
surface hydroxyls, 16-17

Halloysite, 58
Hamaker constant, 211-12
model equation, 212
retarded, 212
H-curve isotherm, 116-17
HSAB principle, 129
Humic substance, 9
adsorption mechanisms, 143-47
functional groups, 18-19
metal adsorption, 135
Hydrogen honding, 14~

INDEX

lAP (ion activity product), 124-27


Illitic mica, 6
metal adsorption, 133-34
specific surface area, 30, 34
Inner potential, 90-91,154-55,206
model dependence, 93
Poisson equation, 95
Poisson-Boltzmann equation, 154
Inner-sphere complex, 13,15-16,145,
178

adsorption selectivity, 129


defined, 13, 178
effect on coagulation, 222
surface charge, 79, 189
Interparticle forces, 205-17
total potential energy, 218-20
Intrinsic surface charge, 36
Ionic potential, 28,63,71,137-38
and interlayer hydration, 28
and organic matter adsorption, 146
and surface protonation, 71
Ionic radii, 3
Isoelectric point, 81,110
I-structure, 48,52
Kaolinite, 6-7
metal adsorption, 18, 133-34
PZC, 84
surface area, 24,30
surface charge, 40
surface complex, 18
surface hydroxyls, 17-18
water structure, 58-61
weathered, 22
Kurbatov plot, 135
Langmuir equation, 27, 118, 124
phosphate precipitation, 124
two-surface, 118
Layer charge, 6,8,38-39,61
L-curve isotherm, 116-17
Lewis acid, 17
Lewis acid site, 17-18
Lewis acid softness, 71-72,129-33,146
relation to cation exchange, 129-33
relation to organic matter
adsorption, 146
Lewis base, 14, 132
Ligand effects on metal
adsorption, 132-38
Ligand-like adsorption, 136, 139, 146,
197
Ligand exchange, 138-41, 143, 145-47
defined, 13H

233

INDEX
evidencefor, 139-41
organic matter adsorption, 147
surface complex model, 165,169
Liquid water, 54
dynamic structure, 48-49

Metal adsorption, 128-38


ligand effects, 132
selectivity, 129
Metal-like adsorption, 136, 139, 197
Misono softness parameter, 71,130
defined, 130
relation to adsorption, 130-31,146
Montmorillonite, 6-8
electrophoretic mobility, 99
metal adsorption, 129-33,135
quasicrystal, 34-35,198-205
surface area, 24-25,30,34
surface charge, 37
surface coating, 21
surface complex, 15
water structure on, 66

Navier-Stokes equation, 95
Negative adsorption, 31, 106
Babcock model, 109
DDLtheory, 107
Neutron scattering, 51, 199, 221

Objective model, 185-88


Organic matter adsorption, 143-47
mechanisms, 143
Outer-sphere complex, 13, 15, 18, 178
defined, 13, 178
effect on coagulation, 223
and metal adsorption, 132
and oxyanion adsorption, 139
quasicrystal, 199
surface charge density, 79,189

Packing area, 26-27, 29


CPB, 28
nitrogen, 27
water, 27
pH effect, metal adsorption, 135,175,
183
organic matter adsorption, 146
oxyanion adsorption, 140
pH,( 135, 137
Phyllosilicates, 4
groups, 6
intentratlflclltion, 111-21

layer charge, 6,8


layer types, 6-8
surface area, 30,34
surface complexes, 15, 18,21
water structure, 57-69
weathered, 22
Point of zero charge, 81, 83-84, 86
Poisson equation, 95
Poisson-Boltzmann equation, 154
analytical solutions, 158
assumptions, 154-55
asymptotic solution, 158,208
integrated form, 156, 207
Polarizable interface, 92
Polymer bridging, 223-24
Potential-determining ion, 93
Precipitation, 122
adsorption versus, 122-28
ion activity product, 124
kinetics, 128
Proton surface charge, 39
measurement, 41-42
PZC, 81,83,86
electrokinetic measurement, 98
mixtures, 87
oxyanion adsorption, 85,139
PZNC, 81,83
relation to PZC, 82, 86
relation to PZNPC, 86
PZNPC, 81,83
relation to PZNC, 86
PZSE, 41,83
relation to PZC, 83,86
shift from adsorption, 85

Quasicrystal, 22,24-25,34,100,198-205
defined, 24, 198
enthalpy effect, 204-5
equilibrium structure, 198-204
formation, 204-5
neutron scattering data, 200-201
smectite, 198-199

Relative surface excess, 106,113-15


Reversible interface, 91
Rotational correlation times, 49

Schulze-Hardy rule, 218


S-curve isotherm, 116-17
Sedimentation potential, 104
Selectivity sequence, 129
Selfdiftlcusion coefficient, 49, S3, 56,
65-67

234

Sheet structure, dioctahedral, 2-3


tetrahedral, 2-3
trioctahedral, 3
Silanol group, 18,40
Siloxane ditrigonal cavity, 13-16
cation exchange, 129
Lewis base, 14-16
organic matter reactions, 146
quasicrystal, 198-99
weathering sequence, 19
Smectite, 6
interstratification, 20
quasicrystal, 198
surface coating, 21
weathered, 22
Solvation force, 213-17
measurements, 215-16
swelling pressure, 214-15
Solvation shell, 54,56-57,214
bivalent ions, 56,62
monovalent ions, 55,67
Sorption, 122
Specific adsorption, 79,85
effect on PZSE, 85
Specific surface area, 23-25
chloride exclusion, 34
CPB adsorption, 29-30
negative adsorption methods, 29
nitrogen adsorption, 28,30,34
phyllosilicates, 30, 34
physical methods, 23
positive adsorption methods, 25
water adsorption, 28-30
Streaming potential, 102
Structural surface charge, 37
heterogeneity, 38-39
Structure, 1-12, 47
allophane, 9
amorphous, 1
crystalline, 1
humic substances, 9-12
oxides, 3-5
phyllosilicates, 4,6-8
sheet, 2-3
water, 47-54
Surface charge density, 35-42
alkylammonium method, 38
balance law, 79
dissociated, 79
inner-sphere complex, 79
intrinsic, 36
outer-sphere complex, 79
particle, 79
proton, 39, 189
structural, 37,86

INDEX

Surface complex, 13, 15, 16, 18


and specific surface area, 34
Surface complexation model, 162-69
activity coefficients, 192
reactions, 165, 190
structure, 188-93
Surface functional group, 12
Lewis acid site, 17-18
orangic, 18-19
oxide, 16-18
siloxane sheet, 13-16
weathering sequence, 19
Surface hydroxyl, 16-19
Surface protonation, 39-42,71-72,
138-139,143-144
and adsorption, 144
model calculations, 170, 179
Surface speciation, SO, 163, 169-85
Swelling pressure, 208,214-15
Tactoid, 221
Triple layer model, 177-85
anion adsorption, 184-85
metal adsorption, 182-84
protonation, 179-82
van Bernmelen-Freundlich
equation, 120-22
van der Waals interactions, 145-47,
209-13
interparticIeforce, 210-11
measurements, 213
van der Waals model, 164-69
conditional equilibrium constant, 167
equilibrium constant, 168
rational activity coefficients, 169, 192
Vermiculite, 6-8
layer charge, 6,61
surface area, 30
surface coating, 21
surface complex, 15
water structure on, 61
V-structure, 48,52-53
Water bridging, 143-45
Water structure, 47-54
D-structure, 54
electrolyte solutions, 54-57
experimental methods, 49
I-structure, 52
V-structure, 52-54
Zeta potential,

96

You might also like