Garrison Sposito - The Surface Chemistry of Soils (1984)
Garrison Sposito - The Surface Chemistry of Soils (1984)
Garrison Sposito - The Surface Chemistry of Soils (1984)
Garrison Sposito
THE SURFACE CHEMISTRY
OF SOILS
THE SURFACE CHEMISTRY
OF SOILS
Garrison Sposito
University of California, Riverside
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 encourage-
ment 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.
IG. N, Lewis und M. Rundall, Thermodynamics and till' Free f.'Tll'fNY of Chemical Substances.
Mc(irllw·llill Book ('0" New York. 11)2,1 Reprinted with the permission or the publisher.
CONTENTS
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 3
Table 1.1. Ionic radii of elements commonly found in clay fractions of soils
Table 1.2. Metal oxides, oxyhydroxides, and hydroxides commonly found in clay
fractions of soils
0/
Fe 3+ Ii ~cJ-'" J
'Xi-p;" d
!~ 9
,1,,"'i
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 com-
bined, 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, vermicu-
lite, and smectite (montmorillonite), each with the general unit cell
chemical formula.f
Chemical formula
coefficients
Layer Layer
Group type charge a b c c'
Kaolinite 1:1 <0.01
Mica (illitic) 2:1 1.4-2.0 6.8 3 c + c' = 0.25
Vermiculite 2:1 1.2-1.8 7 3 0.5
Smectite" 2: 1 0.5-1.2 8 3.2 c + c' = 0.2
Chlorite 2: ) with Variable 2.4 8.4 o.s ).5
hydroxide
intcrlayer
• "nnrl,'lIl1y IIIClllllllllnlllln,I,· III ",II rillY'
_--M m +
0 2-
-<:):~~~X~~~X~--Si4+
I: I LAYER
2:1 LAYER
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 precipi-
tated silica or metal hydroxides in soils, since these compounds typically
are amorphous. Structurally disordered aluminosilicates, known collec-
tively 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 poly-
morph 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. Collec-
tively, 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 decomposi-
tion 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 fun-
damental molecular units are amino acids, which combine according to the
reaction
DEGRADATION OF LIGNIN DEGRADAT ION OF PHENOLS BY
(CONIFEROUS LIGNIN) PROTEINS MICROBIAL SYNTHESIS
ALIPHATIC
CARBON
I SOURCE
o PROTEINS----......
lY OCH PEPTIDES----...... f
y 3
AMINO ACIDS ...............
AMMONIA Rh
COOH
OH
C-
I
-CH RyY0H
~YOH
I
CH20H HO--y /
R~ OH . .
oA-(
OH
RyOH
I
~~ OH
OH
OH
XOCH 3
YCH
OTHER PHENOLS
II
CH OF PLANTS
I
COOH
•
OH
DEGRADAT ION OF
CARBOHYDRATES
THE REACTIVE SOLID SURFACES IN SOILS 11
H
I II °H
I II °
H-N-C-C +H-N-C-C
I I 'OH I I 'OH
H R H R1
°
H I H RI
1
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 )
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-con-
taining 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~£tL£9.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 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 sepa-
rate 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.
THE INORGANIC HYDROXYL GROUP. 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.
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
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 co-
ordinated 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 hydrox-
ide anions. as opposed to the OH groups coordinated to Al(III), which
18 THE SURFACE CHEMISTRY OF SOilS
H 20
LEWIS
ACID SITE
ALUMINOL
siloxane surfaces that permit stacking of the 2: 1 layers along the crystallo-
graphic 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 silox-
ane 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 condi-
tions 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 trans-
formation 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 inner-
sphere 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, mont-
morillonite 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 com-
plexes with K I is inhihited completely and a transformation of the layer
THE REACTIVE SOLID SURFACES IN SOILS 21
/
mica layer oroanlc coat Ino
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.
--
tion. In both techniques. the objective is to determine the shapes and
dimensions of representative soil particles and then to calculate the
24 THE SURFACE CHEMISTRY OF SOILS
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 crystallog-
raphic specific surface area of an Avogadro number of unit cells of mont-
morillonite now can be calculated:
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 mOiecurar··i.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)
r
Table 1.6. Packing areas of gas-phase adsorbates used in specific surface area
measurements
Range of
Gas-phase observed
adsorbate T,K aid
m- nrrr' am, nm 2 am, nm 2
Nitrogen 78 0.162 0.162 0.13-0.20
Argon 77 0.138 0.167 0.13-0.18
Krypton 78 0.152 0.202 0.17-0.22
Ethylene glycol
(1,2-ethanediol) 293 0.224 0.332 0.230-0.332
Ethylene glycol monoethyl ether
(2-ethoxy-ethanol) 293 0.323 0.523 0.396-0.600
Water 293 0.105 0.106 0.075-0.195
BET PLOT
H20 Adsorption
Otago podzol
xm=1/(23.4+ 1.74) =
0.0398kg H20/kg soil
oL-. L--_ _--l.--L_ _--J oL--_---l_ _--l._ _--L_ _--J
Kaolinite (Na-
form, Peerless)" 1.86 1.88
Kaolinite (Bath,
South Carolina? 1.8 1.5
Illitic mica (Na-
form, Illinois)" 10.2
Illitic mica (Na-
form, Fithian, Illinois)" 9.3 9.6
Vermiculite (Mg-form,
llano, Texas)" 0.31 71.2
Vermiculite (Kenya)" <0.1 72.6
Montmorillonite (Na-
form, Wyoming)" 3.3 84.7
Montmorillonite (Na-
form, Wyoming)' 1.4 80.0
a A. G. Keenan et a!., J. Phys. Colloid Chern. 55: 1462 (1951).
b D. J. Greenland and J. P. Quirk, J. Soil Sci. 15: 178 (1964).
C H. D. Orchiston, Soil Sci. 78:463 (1954).
d H. van Olphen, Proc. Int. Clay Cant 1969 1:649 (1969).
e R. W. Mooney et a!., J. Am. Chern. Soc. 4: 1367 (1952).
fD. 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.
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 approxi-
mated 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
-, = (2RTC/F.f1[) 1/ 2 cxp(-Fl/J/2RT) (X = 8) (1.14)
(X
32 THE SURFACE CHEMISTRY OF SOILS
where i3 is the distance between two planes: that where l/J = - 00 and that
where u" is evaluated. The solution of Eq. 1.14 is
exp(Fl/J,,/2RT) = F(2c/ eoDRT)1/2 i3/2 (1.15)
The definition of dex(c) in DDL theory is38
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 dl/J
1 {O
(1 - eY)dy
= ~ )Y3 [c(e- Y + eY - 2)]1/2
= -1- iO ey/ 2 dy
jj3C Y.
-
2
= r;;: [1 - exp(y,,/2)]
v f3c
2
= - - i3 (1.17)
jj3C
where
f3 == 2F 2/e oDRT = 1.084 X 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 dm·l/2kg - I.
THE REACTIVE SOLID SURFACES IN SOILS 33
No-MONTMORILLONITE IN NoCI
10kg CLAY/ m3 SUSPENSION
-
)(
>0)
Vex = 0.5524 + 0.3046c- 1/ 2
r 2 = 0.968
0'---------4-------'-------'--------'
o 10 30 40
Therefore,
SE =~ x (slope/2)
(1.084)1/2 1/2 1/2 10- 3 / 2 m 3 / 2
= x 10 m
8 mol- x ----;;-;",.--
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 pre-
viously. 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
Vex = f f(
1- 1- ~) lSodex(C) + ~ Sod ( 1.19)
THE REACTIVE SOLID SURFACES IN SOilS 35
(1.20)
Uo
5
= (2N
qF)
(10
18
)
Aab
1.0 ....L...L..........,
15 20 0,6 0.7
n LAYER CHARGE
c
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 distribu-
tion of layer charge is plotted in Fig. 1.14. Similar kinds of charge
partitioning involving combinations of X-ray diffraction data with theore-
tical analysis are possible for vermiculitic and mixed-phyllosilicate soil
clays."?
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., inor-
ganic 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 support-
ing 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
qOH/S,
Mineral p,molcm -z Reactive hydroxyl groups assumed"
Goethite 4.4 6.7 Type A OH and Lewis acid OHz
Gibbsite 2.8 5.6 Edge-surface Lewis acid OH and OHz
Kaolinite 0.35 1.0 Edge-surface silanol and aluminol,
Lewis acid OH z
a 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
(1.25)
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
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 Environ-
ments (J. B. Dixon and S. B. Weed, eds.) Soil Science Society of America,
Madison, Wis., 1977. U. Schwertmann, D. G. Schulze, and E. Murad, Iden-
tification offerrihydrite in soils by dissolution kinetics, differential X-ray diffrac-
tion, 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 environ-
ments in South Africa, Geoderma 27:33; (I \1H2).
THE REACTIVE SOLID SURFACES IN SOILS 43
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 mat-
ter 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 montmorillo-
nites, 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). Determina-
tion 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 layer-
charge 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 vermicu-
lites, Clays and Clay Minerals 30: 215 (19H2).
46 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 mole-
cules 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 sur-
rounding 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
Molecular parameter and
Method its physical significance
Infrared and Raman OR and hydrogen bond strength, orientation,
spectroscopy and length (V structure)
Electron spin resonance Solvation water molecule orientation Tc ,
correlation time for rotation of a solvation
complex
Incoherent neutron Ds , self-diffusion coefficient
scattering TR ~ residence time for jump diffusion
TJ, correlation time for rotation of the dipole
moment
Nuclear magnetic resonance 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
Dielectric relaxation TJ, correlation time for rotation of the dipole
moment
a, measure of the spread of correlation times
about TI
Neutron, electron, and Water-molecule 0 and H positions and bond
X-ray diffraction orientations (0 structure)
50 THE SURFACE CHEMISTRY OF SOILS
J r 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)
this fact is that the local environment of a water molecule in the liquid
undergoes many fluctuations during the course of an elementary "single-
particle" motion through translation or rotation, Therefore, the param-
eters 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 hydrogen-
bonded 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 pro-
jected 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] prop-
erties 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."
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 concentra-
tion and temperature of the solution, and with the pressure applied to it.
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"
Figure 2.3. The D structure of water in the interlayers of lO-A halloysite. (After
Hendricks and Jefferson 23 )
1
I
I
I
I ,
- -M---'--'t)-e----
~ ~------ -\ -----1---
\ I
\ I
\ I
I I
60 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 monova-
lent 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, Mg-
vermiculite 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
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 struc-
tures of the solvating water molecules in monovalent and bivalent cation-
saturated 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 two-
layer 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 coef-
ficients 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 geometri-
cally with the D structure in Fig. 2.4. The effective self-diffusion coefficient
for the water protons in the region between solvation shells was deter-
mined 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 self-
diffusion 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
(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 mont-
morillonite 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 mono-
layer of watcr are around 5 ps, according to dielectric relaxation
68 THE SURFACE CHEMISTRY OF SOilS
CK = {MOH(HzO)~~11)+}{NHt} = {NHt}z
(2.5)
(M(H zO):+)(NH3 ) (HzO)(NH3 )
Ionic Misono
Exchangeable potential," softness, 50 Water {NHt}, cK,
cation nm"" nm activity mol, kg" mof
Li+ 13.5 0.053 0.20 0.23 1.32b
0.98 0.17 0.15
Na+ 9.8 0.111 0.20 0.16 0.64
0.98 0.10 0.05
K+ 7.3 0.189 0.20 0.10 0.25
0.98 0.11 0.06
Ca2+ 20.0 0.165 0.20 0.80 16.0
0.98 0.16 0.13
Mg2 + 27.8 0.096 0.20 1.01 25.5
0.98 0.74 2.79
• Calcuhued u.11I1l dulU from 1'1I"le I. I.
hAil vlllue. culeulntcd f"r I kll ",' cllly under the 1I••umpuon Ihlll (NII.l ~ 0.2 III 110,7% .,,11111011 of
NII.lIII
72 THE SURFACE CHEMISTRY OF SOilS
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
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. Hen-
dricks 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 thermody-
namics 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 comprehen-
sively 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
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
.J..
"4 C Poi .
omsrgnon . ~2
I op, cit.'
et 1\.,
THE STRUCTURE OF WATER NEAR CLAY MINERAL SURFACES 77
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 proton-
ated 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 ex-
posed on soil clay particles.
With these additional definitions, the surface density of net total particle
charge can be expressed mathematically:
(Tp => (Tin + (TIS + (TOS
(3.2)
= (TO + (TH + (TIS + (Tos
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:
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 well-
known 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:
F(AEC - CEC)
O"IS = -O"in + S (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
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
F(CEC - AEC)
ars + aos + aD = 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 crystallin-
ity, 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
I
> I
c>
rVI .:<:
u
(\J o
E 0 E
~
E
ex> pH
0
>< -I
::l GOETHITE
Y- AI203
10-3M NaCI 0.015 M NaCI04
HUANG AND STUMM (1973) BAR-YOSEF et al. (1975)
-2
-60
-20 rc> 40
.:<: KAOLINITE
0 u FERRIS AND JEPSON (1975)
0
E
E
I
c> u
.:<: 0"
u HYDROXYAPATITE >-
0
o BELL et ol. (l973l 0
E Z 0
E 0"
KCI pH
'" 1M
o O.IM
POINTS OF ZERO CHARGE
• O.OIM
11.0
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
84 THE SURFACE CHEMISTRY OF SOILS
Table 3.2. Comparison of points of zero charge for several solid phases suspended
in solutions of 1 : 1 electrolytes
Figure 3.2. The downward shift of the PZSE for an Oxisol soil (Typic Torrox) in
response to the specific adsorption of o-phosphate."
:r
0 ~
cr
I
:r
• 1M
0 0.1 I!!
cr
• O.OIM
0 0.001 I!!
86 THE SURFACE CHEMISTRY OF SOilS
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
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 combina-
tion of these two equations then produces the expression
F(CEC - AEC)
0"0 +=0 (3.13)
S
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 dissolv-
ing 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
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 relation-
ship 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.
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 conceptualiza-
tion 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 electro-
chemical 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
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 unambigu-
ously, 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 experi-
mental 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
(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
Polarizable No
Rt oo
Reversible Yes
THE ELECTRIFIED INTERFACE IN SOILS 93
~!~~~~~~~~~~~~~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 thermodyna-
mics of irreversible processes, with no appeal made to specific models of
the interfacial region."? With this approach, of course, detailed informa-
tion 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 thick-
ness of the interfacial region extending into the soil solution. A perfectly
flat clay particle surface meets this condition exactly, for example, be-
cause 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 dis-
tortion 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
SOLID
SURFACE
z
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
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:
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)
dV')
and ( -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
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
96 THE SURFACE CHEMISTRY OF SOILS
=u Ir p(x)dxdy (3.31)
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) == dZIjIz dx
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
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
[b e D(
-~
ffba' [f(x) dPJ (3.34)
U f Ja p(x)dxdy = - dz dxdy
;;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
f(x) = p(x)E -=0
dz
and Eq. 3.34 reduces to
[b e D( I[b
U
f
Jd p(x)dxdy = -~ E Jd p(x)dxdy
or, after cancelling the integral from both sides,
U = 6 oD(E (3.36)
T1
98 THE SURFACE CHEMISTRY OF SOilS
(3.38)
where
+1
= { -1
if c> 0
sgn«()
if « 0
n
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:
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 electro-
phoretic 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
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
u, = -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-mont-
morillonite 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
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
.iVst = KTf (.iP (3.43)
the result with Eq. 3.42, one can derive the symmetric relationship
r; Q
_.::..:..--=- (3.44)
dl'[tiz E
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 elec-
trophoretic 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
KlaVstl/aP = 10- C'm-1 1s- 1V- 1 7
x 10- V'm 2N- = 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 concentra-
tion 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 dC dV
o= L y dz + L YD dz + LYE dz
dP dC dV
o= LEY -
dz
+ LED -
dz
+ LE -
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.
(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 under-
going 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.
p(x)E
d (T/ dVl)
+ dx dx = 0 (3.49a)
With the help of Eq. 3.26 this expression takes the form
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 de,
EeoD-
/ dx
= T/-
dx
(3.49c)
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
(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 proto-
typical 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,
= - 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"
(3.55)
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.. Equa-
tion 3.56 can be developed further in DDL theory as indicated follow-
ing 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 two-
chamber.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)
mM 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 assump-
tion 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:
su su uoSm s (
mNa = mCi - FM 36Oa)
w
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 109
mSu mSu -
uosms) = m 2 (3.61)
Cl ( Cl FMw
11m == S msr Cl
(w )
-uoSm s
Q== FMw (3.63)
Mw
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,
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 Physi-
cochemical 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
(4.1)
114 THE SURFACE CHEMISTRY OF SOilS
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 accumula-
tion. Also note that there is no adsorption of i when the condition
n, Xi
-=- (4.2)
nj Xj
(4.3)
(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)
5
L -curve
S-curve
-
I
01
-'"
I
01 "0
-'" E Anderson sandy
"0 E cloy loom
E Q. pH 6.2 25°C
E 0"
::>
Altamont cloy loom r = 0.02M.
u pH 5.1 25°C
0"
r = O.OIM
00
CUT (rnrnol
H-curve
0.60 150
I
C-curve
01
-'"
"0 0.40 I
01 100
E Boomer loom -'"
E
pH 7.0 25°C 0
Har-Barqan cloy
"0
u E
0" 0.20 Il':l0.005M. ::l.. 50 parathion adsorption
~
0"
from hexane
50% RH hydration
0
0 0.05 0.10 0.15 0.20 0.25 40
Cdr (mmol m- 3)
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
bKc
q = 1 + Kc (4.7)
(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)
II ~ to + "'''
Thus a graph of K d ~ainst q s~ould be 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:"
blKlc b
q- + 2K2c (4.10)
1 + Klc 1 + K 2c
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,
--=
(K. + K 2)Kd + 2K.K 2 q - bK tK 2
(4.12)
dq 2K.. + (K. + K 2)q - (b.K. + b 2K 2 )
INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOILS 119
1.0
(
P SORPTION
WATTS SOIL
t = 25 ± I·C
~ 0.8
o
II>
C'
.><
<, 0.6
c:
o
..-
:;,
o
II> 0.4
'"E
"0
INTERCEPT = a~1 I a II
~ 0.2
D D INTERCEPT=l3o
SLOPE =fJO 1"'1
2 4 6 8 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.
Fig. 4.2, then the limiting slopes and the two x intercepts can be deter-
mined 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 =
f
OO
-00
()
m y
exp(y)c
1 + exp(y)c
dy
where y = In K and m(y) is the weighting factor for the Langmuir term in
(4.16)
In (A/b)
Ym = f3 (4.20)
INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOilS 121
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
I
/ " ,
x {3 = 0.8
\
I \
/ \
t \ {3-05
......................._....:.;-;.;1 .... ....,.,.:.:::.:-...:.......
................ ". -- ---_ .
-'I '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.
I
critical reading of the abundant literature on sorption studies, as well as in
the design of experiments on surface phenomena in soils. ,I
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.
INORGANIC AND ORGANIC SOLUTE ADSORPTION IN SOilS 123
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)
T
qpo. = b po . ( 1 - AI V
nAI
w) (4.25a)
The use of cKso to eliminate AIT from Eq. 4.25a and the additional
definition
(4.26)
(4.25b)
Finally, since
124 THE SURFACE CHEMISTRY OF SOILS
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
(4.25c)
QP04 = 1 + (1/ K 4T
) = 1 / KP0 4T
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 estimat-
ing 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 phos-
phate-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,
(Mm+)a (Ll-t
(4.28)
K so = (MaLb(s))
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
K = (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
(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)
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.
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 spec-
troscopy, which give direct evidence for the presence of adsorbed spe-
cies. 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 precipita-
tion. 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 super-
saturation 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 thF-
phosphate-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 phos-
phate 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
128 THE SURFACE CHEMISTRY OF SOILS
where
exp(k 2 qc)
to = k - tc (tc > 0)
1k2
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
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
Li+ - M+ EXCHANGE Cs+
ON MONTMORILLONITE •
10 •
0
Camp Berteau
Wyoming
Rb+
10
0
E 8
•
J
-
~ 'I;
'i,i,"
+ 6
'I
~
~'I
t ':,/j
+ 0
.~
I
-'"
...J
OQl
4
•
K+
' ,~
II
i
<.!l 'I
<J 0
',I
I 'I
I
2 I
0 ',I
Na+ )
0
•
0 ',I
0 0.05 0.10 0.15 0.20 0.25
YM+ - YL1+(nm)
INOKGANIC AND ()K(;ANIC SOllJ II ADSOKI'1I0N IN SOli S III
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-ligand-
surface 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 consis-
tent with at least one category of ligand effecta" On kaolinite, a sharp
D -- K d C -
_/sorb
-- (4.41)
s Isoln
a-FeOOH
!AT =10-4M
O.IM KN03
120r-----,--,..-----,--,---,
Cu Pb Co
...
<,
'0
E· 80
::l..
~
u
c d
0=-3---:-----:----:---::-------:
8
pH pH
Figure 4.6. The effect of increasing OH- activity on the adsorption of metal
cations by solid soil constituents."
dflOrb) = b (4.42)
( dpH pH-pH", 4
136 THE SURFACE CHEMISTRY OF SOILS
or
Isorb = {l + exp[ - b(pH - pHSO)]}-l (4.43b)
4 Rutile in Co (N03)2
Fuerstenou et 01. (198()
Car
1.67 x 10- 3 M
-
I
2
>
I 0
l/)
C\J
-°
00
0
E
1.67x1O:-5M
x
:J
-I
4 6 7 8 9 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
2000-
~
10'
COO
PhOSPhat~ 0
oX a-FeOOH 00
0
0 150 0
E 0
E 00
"0 0
Q) 0
.a 100 (ArSenate
L-
0
(/)
AI(OH)3(am)
"0
0 MA A At:A A
c:
0 50 A
-
c: Borate~
<t ~
Fe(OH)3(am) -
0
4 6
pH
8 10 - 12
(4.45)
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 particu-
larly 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
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 Chem-
istry 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 iso-
therms and its use in diagnosis of adsorption mechanisms and in measure-
ment 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 treat-
ment and classification of the solute adsorption isotherm. Part II: Experi-
mental 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
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, sesquiox-
ides, and soil materials, Soil Sci. Soc. Am. J. 32: 364 (1968). M. A. Anderson,
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).
57. D. T. Malotky and M. A. Anderson, The adsorption of the potential deter-
mining 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).
58. 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.
59. 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).
60. 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).
61. 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).
62. 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 Clay-
Polymer 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.
(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:
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 adsorp-
tion 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!
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 (5.4a)
dx dx
and Eq, 5.3 can be used to convert Eq. 5,1 to the differential equation
dl/J u(x)
-=-- (5,5)
dx enD
156 THE SURFACE CHEMISTRY OF SOILS
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] (5.6)
( -d
x eoD i
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
+1 l/J>O
sgn( l/J) = { -1 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
dx = -sgn(l/J) eoD~ ci[exp(-ZiFl/J(X)/RT) -1]
}1/ (5.8)
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) = L CiZiF exp( -ZiFl/J(X)/RT) (5.9)
i
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
CHEMICAL MODELS OF SURFACE COMPLEXATION 157
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+ = Co = c:
dy
dx = -sgn(y)K[exp( -Zy) + exp(Zy) - 2r/2
= .!.In[tanh(ZY(X)/4)]
Z tanh (ZYo/4)
= -K J: dx' = -KX
where
2 e" - e- u
cschu= U U tanh u = U -u
e - e e +e
and Yo ;& yeO). Upon rearranging this result to be an equation for ",(x), one
obtains
. (1 + 2eYO)1/2 + J3
With b' = (1 + 2eYO)1/2 _ J3' Yo = Fl/J(O)jRT
6c exp(J3K'X) }
1:2 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
has been used. The inverse hyperbolic tangent has the MacLaurin expan-
sion
u3
tanh-1u = u + - + ...
3
which, for large values of the distance x (i.e., ZKX ~ 1), permits Eq. 5.12
to be replaced by the approximation
RT
r/1(x) == 4 ZF a exp( - ZKX) (5.14)
Equation 5.14 shows that the DOL inner potential decreases exponentially
CHEMICAL MODELS OF SURFACE COMPLEXATION 159
according to Eq. 5.7. The special case of Eq. 5.15 that occurs when x is the
position of the electrokinetic plane of shear (and ",(x) = n
was presented
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 a exp( -KX)
O'o(x) = f3 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 smec-
tite siloxane surface. Then K = J f3co = 1.04 X 109 m- 1 at 298 K, a =
tanh (-2) = -0.96403, and .
8(9.6487)(1.04) 10- 3 x 0.96403
O'D(O) = 1.084 x 1 - (0.96403)2
2
- 1.01 C·m
160 THE SURFACE CHEMISTRY OF SOilS
u
0
<,
I
3~ \ Co =100 mol m- 3
CT =-0.266C m- 2
0
o
C\J
<,
3
CATIONS
0
0
\ c 0 =50molm- 3
CT =-0.177C m- 2
I 0
Q.. 2 CATIONS Q... 2
0
~ ~
0 0
u
0 0
o
<,
I 0
0
<,
I
~
+
Q.. ·1 ANIONS. •
0' ~ I I I I , I I
5 10 15 20 o 2 3 4
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 mu-
tually 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."
(5.22)
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(sys-
tern) 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 + pZp(aq) (5.25a)
SRZSR(S) + qL1-(aq) + xH+(aq) = SHxL~sC + 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 com-
plexation 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
SRzsR Mm+ Ll- pZp Rr - Reaction
SoO- a Na+ - - - SoO-(s) + Na+(aq) = SoONa(s)
SOH b Na+ - H+ - SOH(s) + Na+(aq) = SONa(s) + H+(aq)
SOH - Cl- - - SOH(s) + Cqaq) + H+(aq) = SOHzCI(s)
SOH Cu2+ FUL- C H+ - SOH(s) + Cu2+(aq) + FUL"(aq) = SOCuFUL(s) + H+(aq)
SOH Pb z+ - H+ - SOH(s) + Pbz+(aq) + OH-(aq) = SOPbOH(s) + H''{aq)
SOH P0,43- - OH- SOH(s) + POl-(aq) + 2H+(aq) = SHZP04(s) + OlF'(aq)
SOH - Cit3- d - OH- SOH(s) + Cit3-(aq) + H+(aq) = SHCiC(s) = OH-(aq)
• Siloxane ditrigonal cavity.
b Inorganic surface hydroxyl group.
e Fulvic acid anion.
d Citrate (2-hydroxypropane-l ,2,3-tricarboxylate).
(5.27)
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 _ QSR'Cexp[(JLc - ZSR'C e t/J~V;:~/ kBT](pzp)
K - (5.29)
QSR exp[(/LR - ZSR e t/J~i")/kBT](C)
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)
168 THE SURFACE CHEMISTRY OF SOilS
where
KSR'c = QSR'C exp(p,o[C(aq)]/ kBT)
(5.32)
K SR = QSR exp(JLO[pZp(aq)]/ kBT)
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
In K = fa! In -« dXSR'C
for the thertllodynamic equilibrium constant pertaining to Eq. 5.25a shows
that, in the van der Waals model,
K = (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
CHEMICAL MODELS OF SURFACE COMPLEXATION 169
fSR'c K 0 0
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.
where
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 func-
tion in Eq. 5.15 by its first three terms in a MacLaurin expansion.
THE NET PROTON CHARGE. 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)
SOH(s) = SO-(s) + H''{aq) (5.41b)
The conditional equilibrium constants for these two reactions are 15
KS = xsoH[H+] (5.42a)
al - +
XSOH 2
(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) = K~2 exp( -Ft/lslRT) (5.43b)
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 UH (5.44a)
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
(5.45)
pH> PZNPC
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.
8
-log K~I (int) = 7.2
_o~
~ r ~
~o. ~ .~ ~
--- . •
0' 5 6- 6- 6- 6- 6- 6- 6- 6-
If
~
0 6-
y-A1 20 3 10.0 6-
I
6-
O.IM NaCI04 -I
9.5~
:::: log
-Kg (int) =9.5
I
2
21 I I I I· I I I I I I
9.0
0 0.08 0.24 0.40 0.56 0.72 0 0.08 0.24 0.40 0.56
XSOH+2 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- ....... -••• --
CHEMICAL MODELS OF SURFACE COMPLEXATION 173
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
S' _ K so - U max
Kdmt) =-- C=--o- (5.47b)
K SO H l/Jso-
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
Figure 5.2, with its two lines of unequal slope (in absolute value),
illustrates another typical feature of applications of the constant capaci-
tance 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.
* s _ {(SOh M(m-2)}[H+]2
f32 = {SOH}2[M m+] (5.51b)
100
o
w Cu(II)
CO
0:::
g 60
o
<{
~
Z •
W
()
0:::
W
Q.
SILICA GEL •
3M NaCI0 4
2 3 4 5 6 7 8 9
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 )
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
(5.53a)
s = {S2HxL~Y+1)}[OH-Y
f32 - {SOHf[HxL~X-ql)] (5.53b)
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
CHEMICAL MODELS OF SURFACE COMPLEXATION 177
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:
(TOS and that surface charge balance (Eq. 3.36) is expressed mathemati-
cally by the equation
(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
xso-(H+)
Qa2 == --=.=---..;'--...:.. (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 xso-
are 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 equilib-
rium constants for these latter reactions are, for the common example of a
1: 1 background electrolyte ML(aq)
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)'
pH -log' , ,
= -logK~~t - e UH - e"'f3 (562a)
XSOHt (In 1O)C1kBT (In 1O)kBT .
SO M e
pH _IOg(x ) + 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
I I r-----,.----r---r-----,-----,-----,-----,----.-..,
Rutile in LiN03
+
...J
o
*o0-
LiT
.. 1M
• O.IM
6
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)
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
(5.68a)
(5.68b)
The triple layer model has been shown to provide quantitative descrip-
tions 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 surface-
induced 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
o
o
60
.""
50- _Pb 2+ and 50- - PbOH+
-...
cCI.l:: 40
u
CI.l
0..
20
--,-- -- ------
_____ - - - - - 50--Pb 2 + only
ANION ADSORPTION. The triple layer model postulates that anions react
with surface hydroxyl groups according to the general equation
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 com-
plexation 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 con-
firmation 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 agree-
ment is likely to be a model artifact since the mechanism of strong
oxyanion adsorption is ligand exchange, as discussed in Sec. 4.4.
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:
constraint
(5.75)
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
aH -
- F(r(w)
SOH~ -
r(w) ) -
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)
j
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 po-
tentials 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
~S = L L L a{(S(i)O)aM;P(OH)yHxL~k)}
A a j,k p,y,k,q
+ LL L b{S~)M;,n(OH)yHxL~k)} (5.82)
b j,k p,y,x,q
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
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)
ao + aH + aIS + aos + aD + 0 (5.86)
This equation and Eq. 5.82 provide the mass and charge balance con-
straints 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(k»q . a
. (S(i)M(j)(OH) H L(k»(OH)b- y
K(') - b Pyx q (5 87b)
sc - (S(i)OH)b(M(j»)p(HY(L(k»q .
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.
where
CK(I) _ xsodH)a
soc - xS'OH(C) (5.89)
fsoc and fSOH are rational activity coefficients.P Given that K~bc can be C
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:
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
layer model. Because the inner potentials are not determinable indepen-
dently, 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 develop-
ment 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
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 double-
layer 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
PHT. . ewan,' op. cit. ·20
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 inter-
face, 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 197
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
(6.4)
h
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 param-
eters 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 aggre-
gate, the concentration of clay in the suspension, and the number of unit
layers in face-to-face association within the aggregate." Under the assump-
tion 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 suspen-
sions 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 electropho-
retic mobility of its constituent particles are expected to be directly sensi-
tive 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
•
0.8
0.6 •o
o TRANSMITTED LIGHT
INTENSITY
• ELECTROPHORETIC MOBILITY
o INTRINSIC VISCOSITY
• CHLORIDE EXCLUSION
o 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 quasi-
crystals 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.
Eq. 5.9. In differential form this charge density obeys the equation
RT
zpdpj = -cjZjFexp(-ZjFl/J/RT)dl/J
I
= -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,
P = RT}2 (pJZjF) (6.6)
i
-d [ P(x) - -eoD
dx
1
2
(dl/J)
-
dx
2] = 0 (6.7b)
iPo
dP = - fc"'m
0
p(l/J)dl/J
where Po is the pressure in the aqueous solution far from the two planes,
(6.8)
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
where all of the symbols in the second term on the right side are as defined
following Eqs. 1.11 and 5.1. This model expression gives the applied
pressure required to maintain the two planar surfaces bearing the charge
density O'p a distance d apart at equilibrium.
SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY 207
dl/J)
l/J(d/2) = l/Jm ( -dx x=d/2 -0 (6.10)
(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 (6.16)
= Po + 64a 2cRT exp( -ZKd)
In this approximation, Pm is an exponentially decreasing function of the
interplane separation. The potenial energy per unit area, 'Pm' correspond-
ing to the force per unit area, Pm' can be calculated readily by a standard
integration: 16
12 2cRT j:12 exp(-2ZKg)dg
'Pm = -2 j: [Pm(g) - Po]dg = -128a
64a 2
= ZK cRT exp(-ZKd) (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 concentra-
tions. At higher concentrations of KN0 3 (above 10 mol, m'-3), however,
SURFACE CHEMICAL ASPECTS OF SOIL COLLOIDAL STABILITY 209
104 .___-.....---,.---.--..,.--"""T""-~-_,_-_..,.-_,--.___-~-_r_...,
MUSCOVITE IN KN0 3
20°C pH6
• O.lmol m- 3
o I mol m- 3
10
o 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
THE VAN DER WAALS DISPERSION FORCE. When considered over a time
16
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
210 THE SURFACE CHEMISTRY OF SOilS
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 aver-
age 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)
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
A
-
121Td 2 (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
o ....-----.-------.----::-........--....-0
-50
-
( \J
IE -100
"'?
--::l o
-
"'C
-ft
~
-150
I::: MUSCOVITE
C\J
• KN0 3
o Co(N0 3)2
-200
o 5 10 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 SOLVATION FORCE. 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 dielec-
tric 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
214 THE SURFACE CHEMISTRY OF SOilS
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
102 FORCES
o 10 20 30 40 50
d(nm)
216 THE SURFACE CHEMISTRY OF SOILS
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"
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 chemi-
cal 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
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
220 THE SURfACE CHEMISTRY OF SOilS
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 a
ZK ccc RT exp( -ZKcd) = 127Td 2 - 127T exp( -dI5) (6.27a)
c
whereas the second condition on ((J(d) yields the expression
A al5
64 a2 ccc RT exp( -ZKcd) = 67Td3 - 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 concen-
tration 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 concentra-
tion 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-,« = 2 (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
a RT) -3
2
.3/ _ (30721T
KC eee - e2 A Z
or
_(30721Ta 2RT)2 -6
eee - 2 3/2 Z
e f3 A (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).
COMPLEXATION REACTIONS. 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 parti-
cles into parallel stacking along a direction perpendicular to their crystal-
lographic 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
• ! £ • I I' •• ..
222 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*