Sie sind auf Seite 1von 18

Chapter 33

Aluminum Hydrolysis Reactions and Products


in Mildly Acidic Aqueous Systems

John D. Hem and Charles E. Roberson

U.S. Geological Survey, Water Resources Division, Mail Stop 427, 345
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

Middlefield Road, Menlo Park, CA 94025


Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

Early stages of the hydrolysis and hydroxide polymerization reactions of


aluminum have been studied using a constant slow aluminum perchlorate flux
into a solution held at constant pH by addition of dilute NaOH from a pH stat,
observing rates of base addition and periodically determining concentrations of
the various species of dissolved aluminum. Total final A l concentrations

generally were between 2 x 10-4 and 5 x 10-4 molal and holding pH's ranged
from 4.75 to 5.20. Polymeric ionic species began to form when supersaturation
with respect to microcrystalline gibbsite reached an affinity of reaction (A) value
near 1.0 kcal at pH 4.75, and the A value remained nearly constant during the
remainder of the titration. Similar behavior, but with somewhat higher A values,
was observed at higher pH's with a maximum A near 1.6 kcal at pH 5.20.
Conditional first order rate constants were determined for the formation of
polymeric ionic units with OH/Αl molar ratios from 2.0 to 2.4. Average log k"
values (sec-1) ranged from -4.59 at pH 4.75 to -3.25 at pH 5.20 for experiments
done at 25°C. Rate constants about 0.6 log units less negative were measured at
35°C, and about 0.9 log units more negative at 10°C. Species whose OH/Al
ratio was greater than 2.4 were formed at the higher pH's, and developed into
microcrystalline gibbsite. Polymeric aluminum hydroxide ions and colloidal
microcrystalline gibbsite may occur in river and lake waters, especially those
affected by low-pH precipitation, but positive identification of such material is
difficult.
An understanding of the mechanisms of aluminum hydrolysis and the formation of crystalline species
of aluminum hydroxide has been viewed as important in various fields of pure and applied chemistry,
biochemistry, and geochemistry. In part, this interest results from the unique properties of certain
hydrolysis species of aluminum that appear to be present as polymeric or polynuclear macro-ions.
These ions have a strong positive charge and may interact with specific charge sites on surfaces they
encounter. The polymeric species also may grow by accretion, and they may persist mctastably for
months or years under some conditions (1).

This chapter not subject to U.S. copyright


Published 1990 American Chemical Society

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
430 CHEMICAL MODELING OF AQUEOUS SYSTEMS II

Aluminum hydrolysis reactions are particularly sensitive to p H and temperature, and to the
concentrations of other ligands that might compete with O H in complex formation. The most
important inorganic competitor is F (2). Certain organic ligands also can form strong complexes with
aluminum (3). Hydrolysis reactions of aluminum also are influenced by sulfate, and hydroxysulfate
solids may control aluminum solubility in acidic solutions (4).
This paper presents results of some open-system laboratory experiments in which aluminum
hydrolysis behavior was studied in detail between p H 4.75 and 5.20, at 1 0 ° , 2 5 ° , and 3 5 ° C . Results
are compared with those of our earlier work and that of others on the hydrolysis of aluminum in dilute
solutions below the p H of minimum aluminum solubility.

P R E V I O U S WORK

In earlier studies of aluminum hydrolysis reactions (1, 5) test solutions in which ΟΗτ/ΑΙτ < 3.0 were
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

prepared by batch-wise mixing of stock solutions of A1(C104)3, HC104, NaC104, and NaOH. The
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

analysis procedure separated the aluminum into three forms based on the kinetics of color
development with ferron (8-hydroxy, 7-iodo, 5-quinoline sulfonic acid). A related technique using 8-
quinolinol was used earlier by others (6-7). More recently, the ferron procedure was used by
Parthasarathy and Buffle (8) and by Bersillon et al. (9) to study aluminum hydrolysis at higher Αΐτ
concentrations. Jardine and Zelazny (10) described in detail the nature of interactions between the
ferron reagent and aluminum species. The three forms of aluminum are designated (1) as: Ala, very
3+ (3-x) +

rapidly reacting, presumably consisting of uncomplexcd A l and monomeric Al(OH)x complexes;


Alb, a form reacting with psuedo - first order kinetics having a half-time of 12 to 20 minutes, and
3 x ) +
presumed to consist of polynuclear A L ^ O H / X " ionic species in which A l ions are bound together by
double O H bridges; and the third form, Ale, that did not react with ferron at a significant rate over a
period of several hours. Ale was shown by electron microscopy to consist of small crystals of
gibbsite.
Smith and Hem (1) concluded that the amount of the Alb polymer formed in their experimental
solutions was, in part, a function of the rate of addition of the NaOH solution during preparation. Fast
addition tended to produce larger proportions of Ale. Slower addition produced less Ale and more
Alb. This behavior is probably related to the higher initial degree of local supersaturation that occurs
during rapid addition of base.

E Q U I L I B R I U M SOLUBILITY OF GIBBSITE

In the absence of complexing agents other than O H equilibrium solubility can be represented as a
3 + 2+
summation of the concentrations of 4 Ala species, A l , A 1 0 H , Al(OH)2, and A l ( O H > . Concentrations
of individual solute species can be derived from thermodynamic activities used in equilibrium
computations by means of the Debye-Huckel equation and appropriate values for ionic strength. It
should be emphasized that the polymeric material, Alb, is a metastablc reaction intermediate, and does
not directly participate in the reversible equilibria controlling gibbsite equilibrium solubility.
Table I contains standard free energies of formation for monomeric aluminum, and aluminum
hydroxide complexes and two forms of crystalline gibbsite. Chemical equilibria and stability data in
Tables II and III can be used with the CAI summation to compute equilibrium aluminum solubility as
3

a function of p H at several temperatures and ionic strengths. The results of such calculations for
microcrystalline gibbsite for the p H range of 4.0 to 6.0 are given in Figure 1. Over the 30 ° C range
from 5.0 °C to 35.0 ° C the solubility at p H 4.00 changes by more than 2 log units, with the highest
solubility at the lowest temperature. The temperature effect is still nearly as great at p H 5.00 but is
much less at p H 6.0, where the solubility at 5.0°C is only about a factor of 2 greater than the value for
25.0 °C.

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
33. H E M & ROBERSON Aluminum Hydrolysis Reactions and Products 431
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
432 CHEMICAL MODELING OF AQUEOUS SYSTEMS II

Table I. Chemical Thermodynamic Data for Aluminum Species

Species Source of data

-117.0 ±0.3 Offi)


-166.9 ±0.3 Calc. from data in (41)
Al(OH)2 -216.6 Calc. from data in (48)
Al(OH> -312.8 Calc. from data in (48)
Al(OH)3C
microcrystalline gibbsite -274.3 ± 0.3 (12)
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

synthetic gibbsite -276.1 (42)


Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

bayerite -275.6 (43)


H2OI - 56.69 (44)
OH" - 37.594 (44)

Another feature indicated by this plot is the displacement of the minimum solubility to lower
pH's as temperature increases. The strong effect of temperature on Al(OH)3 solubility and aluminum
solute speciation was also noted by Seip et al. (Π), who pointed out that the effect could have
environmental significance. Extrapolation of the gibbsite solubility relationship to pH's above 6.0 is
probably not justifiable, as a different form of Al(OH)3c, bayerite, is formed in alkaline solutions (12).
"Microcrystalline gibbsite" is defined here as an Al(OH> solid having the crystalline form of
0
gibbsite and a hydrolysis stability constant of 10 at 25 C, as indicated in Table II. This is based on
experimental data of Hem and Roberson (12) who assigned an uncertainty of ± 0.3 log unit to their
value. Electron micrographs of solids recovered from their solutions by passing them through 0.10
μηι porosity filters showed some hexagonal crystals with maximum dimensions near 0.10 μπι, along
with many much smaller units. Calculations of edge and surface interfacial energies for the smaller
particles (1) showed that the difference in solubility between microcrystalline and macrocrystalline
synthetic gibbsite can be interpreted as a particle-size effect. As noted by Parks (12) the solubility of
such material tends to be controlled by the smallest particles present. Under some conditions of
solution chemistry, microcrystalline gibbsite alters very slowly and material having the solubility
indicated here may persist for more than 3 years of aging (1) in solution at 25 °C. Data in Table II
demonstrate that where solubility is controlled by hydrolysis equilibria this microcrystalline form is
more soluble than well-crystallized gibbsite by a factor of 20.

EXPERIMENTAL PROCEDURES AND RESULTS

The new experimental results that are discussed here were obtained by modified laboratory
procedures. A controlled steady-state hydrolysis at constant ionic strength was accomplished by
adding, with an autotitrator, a dilute solution (4.53 χ 10 molal of Α1(Ο04> at a very slow constant
rate to a reaction vessel containing at the start 150 ml of background electrolyte solution (0.010 molal
NaC104). The solution in the reactor was continuously maintained at a pre-selected pH by a second
autotitrator, operating in a pH-stat mode that added 0.010 molal NaOH through capillary teflon tubing
on demand. Both titrating solutions contained enough of the background electrolyte to bring their
ionic strength to 0.01, and the A1(C104> solution contained enough HC104 to maintain its pH near 3.5.
The reaction vessel (a covered teflon beaker) was immersed in a water bath that held the temperature

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
33. H E M & ROBERSON Aluminum Hydrolysis Reactions and Products 433

constant (±0.1 °C), and the solution was rapidly stirred with a mechanical stirrer. Amounts of
reagents added and rates were recorded continuously on strip charts. A record of amounts of solution
removed was also maintained as samples for Al species analysis were withdrawn using a calibrated
syringe. A glass-calomel combination electrode (Radiometer GK 2401C) was used as pH sensor.
Equipment used included Radiometer ABU80 autoburettes, PHM84 pH meters, TTT 80 titrators, and
REC 80 servographs (use of brand names is for identification only and does not imply endorsement by
the U.S. Geological Survey).

Table II. Equilibrium Constants for Aluminum Hydroxide Precipitation


3+ +
Reaction, A l + 3H 0 t Al(OH>c + 3H ,
2

at Temperatures Indicated
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

Solid species log *K T,°C Source of data


so
Microcryst. gibbsite -10.40 10 a/
Microcryst. gibbsite -9.35 25 (12)
Microcryst. gibbsite -8.72 35 a/
Synthetic gibbsite -8.11 25 (45)
Synthetic gibbsite -8.04 25 Calc. from data in (42)
Synthetic gibbsite -8.04 25 Calc. from data in (46)
Synthetic gibbsite -8.60 15 Calc. from data in (46)
Synthetic gibbsite -7.52 35 Calc. from data in (46)
Synthetic gibbsite -7.97 25 Calc. from data in (47)

a/ Calculated from Van't Hoff equation using values of ΔΗ°Α1(0Η)3 amorph = -305 kcal/mol,
+
ΔΗ°Α1 = -127 kcal/mol, and ΔΗ° mol = -70.41 kcal/mole given by Wagman et al. (44)·

From the titration and analysis data, the concentrations of Ala, Alb, and Ale can be calculated
for any specific time. The concentrations of OH over and above the amount used in reacting with
free H in the aluminum perchlorate titrant solution can be apportioned among these aluminum species
3+ 2+
by assuming that: 1) Ala consists of A l , A10H , Al(OH)2, and Al(OH>, all of which are at
equilibrium with one another at the pH of the solution in the reactor; 2) AL is Al(OH)3c, hence OHc is
three times the determined value of Ale; and 3) the remaining OH is combined with Alb.
Titrations were performed at pH 4.75, 4.90, 5.00, and 5.20 at temperatures of 25°C, and for
some of these pH's titrations were made at 35° and 10°C. Generally, the titration phase of each
experiment was run for a total of 20 to 30 hours, but titration was not continuous. The equipment was
operated for 5 or 6 hours each day for about a week and was shut down overnight except for the
temperature control bath. Samples were taken at the beginning and end of each day's run, and
sometimes more frequently. The final total aluminum concentrations reached in the experimental runs
ranged from about 2.0 χ 10 to about 5.0 χ 10 molal, but in most of the titration period in each
-4 0
experiment the Ala fraction concentration was near or below 10 molal. Characteristically there was
a short period of adjustment at the start of each day's run, after which the rate of base addition
stabilized and remained nearly constant.

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
434 CHEMICAL MODELING OF AQUEOUS SYSTEMS II

Table III. Aluminum Hydrolysis Equilibrium Constants


at Various Temperatures

Reaction T,°C Log Κ Source of data

Al 3 +
+ H 2 O 2 AlOH * + H
2 +
10 -5.44 (48)
25 -5.00 (48)
35 -4.73 (48)
Al 3 +
+ 2H2O t Al(OH) + 2H
2
+
10 -11.2 (48)
25 -10.1 (48)
35 -9.5 (48)
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

3 +
Al + 4H2O t Al(OH)4 + 4H +
10 -24.3 (48)
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

25 -22.7 (48)
35 -21.7 (48)
H2O t
+
H + OH" 10 -14.53 (49)
25 -14.00 (49)
35 -13.68 (49)

Figure 2 is a graph showing the concentrations of Ala and Alb at the end of each day's run
during a total titration time of 1,240 minutes in experiment Al 12, done at pH 4.90 and 25°C. The
tendency for Ala concentration to be maintained at a nearly constant level was observed in all the
titrations. However, the steady-state concentration was somewhat lower at higher pH's. The
concentration of Alb increased at a nearly steady rate during this experiment. No Ale could be
detected in this solution, but some Ale was produced in titrations made at pH 5.00 and 5.20 at 25 °C
and larger amounts of Ale were produced at 35 °C. Detection limits for the procedure are near 10 '
molal.
In the open system maintained during these titrations the hydrolysis reactions are driven by the
continuous addition of Al ions. The rate of the hydrolysis reactions can be inferred from the rate of
OH addition, with appropriate adjustments, and the reaction kinetics can be evaluated. Early in the
titration period the formation of monomeric aluminum hydroxy-ions (Ala) is the only reaction
thermodynamically feasible as the solution is below saturation with respect to microcrystalline
gibbsite. The solid point in Figure 2 represents the equilibrium concentration of Ala with respect to
that solid at 25 °C, I = 0.01 and pH 4.90, which was reached in about 50 minutes.
The second stage, polymerization of monomers to form Alb, did not begin in any of these
experiments until after substantial super-saturation with respect to microcrystalline gibbsite was
reached. Supersaturation is expressed here in terms of reaction affinity, A, in kilocalories (kcal) from
the relationship: A = -2.303RT (log Q - log K) where R is the gas constant, Τ is temperature in the
kelvin scale, Q is the activity quotient, and Κ is the equilibrium constant for the reaction being
considered. Positive values of A are required for the chemical reaction to proceed to the right as
written. The threshold for polymerization to begin is near 1.0 kcal per mole of A l at pH 4.75 and
4.90. Threshold A values were somewhat higher at higher pH.
The third stage in the reaction represents conversion of Alb to Ale by growth of individual units
0
to a size displaying solid-state behavior. At 25 C the critical level for A for Ale formation was near
1.30 kcal. Table IV summarizes the affinity data for all the experiments considered here. The
3+
reaction affinities are affected by pH, temperature and rate of A l addition, and these factors need to
be considered in any interpretation of the data.

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
33. H E M & ROBERSON Aluminum Hydrolysis Reactions and Products 435

Table IV. Reaction Affinities Observed in Titrations at


Various pH's and Temperatures

Reaction affinities, precipitation


of microcrystalline gibbsite, kcal

Experiment pH T,°C 1/ 21 2/

A114* 4.75 25.0 0.89 1.01 0.95


ΑΙΠ* 4.75 35.0 1.13 1.14 1.18
A112* 4.90 25.0 .97 1.20 1.13
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

1.10
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

A115* 4.90 35.0 1.26 1.25


Al 19* 5.00 25.0 1.20 1.34 1.28
A120* 5.00 25.0 1.47 1.50 1.49
A113* 5.20 25.0 1.38 1.28 1.34
Α121 Λ
5.20 25.0 1.62 1.55 1.60
A116^ 5.20 10.0 1.11 1.18 1.15

u Δ at first appearance of Alb. 4/ A l addition rate 0.055


2/ A at end of titration. μιτιοΐε/ππη.
3/ Mean A for titration period. sj A l addition rate 0.122
μπιοΐε/ππη.

The three stages of the hydrolysis reaction can be represented schematically by one
equilibrium and two irreversible reactions as follows:

Al 3+
+ H2O t AlOH * + H 2 +
(1)

A10H + H2O -» 1/6 Al6(OH)S + H


2+ +
(2)

1/6Α1β(ΟΗ)5 + H2O -» Al(OH)3c + H . +


(3)

The sum of reactions, 1, 2, and 3, is: A l + 3H2O t Al(OH)3c + 3H . Values of Κ for the
3+
+

summarizing reaction are given in Table II, and Ki values are given in Table III. A continuous state
of equilibrum for 1 implies that A = 0. Any departure from zero for the observed values for A for the
x

summarizing reaction then can be assigned to the sum of reactions 2 and 3 and is designated A23 .
The formula Α1ό(ΟΗ)ΐ2 represents Alb in equations 2 and 3 above and is equivalent to a single
+
hexagonal ring of 6A1 ions bound together by six (OH)2 bridges. The fundamental distinction
between Ala and Alb species, as earlier studies pointed out, is that the OH in the Alb species is present
in structural bridging positions.

KINETICS OF Alb FORMATION

Data in Table IV for experiments A l 19 and 20 and for A l 13 and 21 indicate the effect of rate of
addition of A l on the reaction affinity. In both pairs of experiments, when A l addition rate was
increased the value of A increased. This difference in A results from the higher steady-state

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
436 CHEMICAL MODELING OF AQUEOUS SYSTEMS II

concentration of Ala that is maintained by the faster rate at which reactants are supplied. The rate of
Alb formation also must increase to maintain the steady-state condition.
In effect this is a self-organized system that maintains a steady state, although not at chemical
equilibrium, by a balance of reaction rates against reactant fluxes, a condition commonly seen in real-
world systems. The titration experiments permit a rather direct measurement of the rate of nucleation
and polymerization reactions that transform Ala monomers to Alb polymers. The titrations at pH 4.75
and 4.90 yielded only Ala and Alb. Although some Ale was formed in titrations at pH 5.00, the
polymeric product was about 90% Alb. Hence, the data obtained in these experiments should be
useful for determining the effect of pH on the rate of the polymerization process leading from Ala to
Alb. At least a part of the information from titrations made at pH 5.20 should also be useful because
during much of those experiments the principal product also was Alb.
The concentration of reaction product, Alb, is assumed not to affect the forward rate of reaction
2 by direct reversal, but it could influence the overall observed rate if Alb rapidly converts to Ale.
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

This process could become coupled to the first step and alter the mechanism. Preliminary titrations
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

made at pH 5.40 produced mostly Ale and the data are not included here.
As the addition of unreacted Al continues, the concentration of Ala should stabilize at a value
governed by the relationship between the rate of formation of Alb and the entering A l flux. The
formation of Alb under these conditions should be a psuedo first-order process whose rate will be
+
indicated by the rate at which OH must be added to react with H produced in the conversion of Ala to
2
Alb. The determined concentration of AlOH * and the properly adjusted rate of OH addition over a
suitable time interval permits calculating a conditional rate constant.
The rate of OH addition taken from the recorder trace of the pH-stat must first be adjusted to
account for the free H that is added with the aluminum perchlorate solution. A further adjustment is
required for the OH taken up by monomeric Ala species, assuming all these species are at
equilibrium, controlled by the titration pH.
The three equilibrium expressions for the formation of monomers are:

2 3+ 1
[AlOH *! = [Al 1 *Ki [ H T (4)

+ 3+ + 2
[Al(OH)2 l = [Al ] *Ki2 [H ]" (5)

3+ + 4
[Al(OH>"l = [Al 1 *KH[H ]" (6)

The concentration of Ala, CAla is represented by the species summation:

CAI. = [Al *] γ" 3 + [A10H ] γ"


3

1
1
+
2+

1
1
2 + + (7)
Al AlOH
+ 1 1
[Al(OH) l γ" 2 • + [Α1(ΟΗ>'] γ"
V 1 V r 1
' Al(OH)2 Al(OH)4
where γ terms are activity coefficients and bracketed quantities are thermodynamic activities.
Concentrations of OHa are computed from

COHa = CA10H 2 +
+ 2CA1(0H)2 + 4CA1(0H)4 . +
(8)

From these equations and data in Table III the value of the molar ratio COHa/CAia for any specified pH
can be computed. Because all the added A l must pass through the monomeric hydrolysis stage, it
follows that the flux, φ AIT, being added to the reactor by the autoburette, can be used in place of C A l a
to calculate an equivalent OH requirement analogous to C O H a .

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
33. H E M & ROBERSON Aluminum Hydrolysis Reactions and Products 437

The remaining OH flux, <j)OHb, can be assigned to production of Alb. The concentration of Alb
is known from the analyses of the solutions. If the ratio COHb/CAlb has a value of 2.0, the
stoichiometry for conversion of Ala to Alb can be represented directly by equation 2, and one mole of
added OHb produces one mole of A l in the form of Alb. If the OHb/Alb ratio is greater or less than
2.0 the rate of OH addition requires a further correction to give a rate constant in terms of aluminum
that is reacting to form Alb. This correction is designated η in the final rate equation, which is
written:

ηφΟΗό = dCAib/dt = k 2 CAIOH 2 +


(9)

and η is computed from the relationship:

1
η = (COHb/CAlb - 1.0)" . (10)
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

2+
A10H is the predominant form of monomeric aluminum over most of the pH range where the
kinetic experiments were made. The OHb flux is converted to moles L sec by dividing the
calculated flux in moles per second by the volume of solution in the reactor at the midpoint of the time
interval being considered. Rate constants were calculated using time intervals of 1 to 2 hours.
Results of the titration experiments are summarized in Table V. Figure 3 is a plot of the
M
conditional rate constant data for these experiments vs the controlled pH. The observed range in k2
for a given pH is ±0.2 log units. There is a clear indication in Figure 3 of a well defined pH
dependence of the nucleation and polymerization rate. The dashed line on the graph represents a third-
order pH dependency - that is, a change of 0.5 unit in pH causes a change in la" of 1.5 log units.

Table V. Typical Titration Data Used for Computing Kinetics of Al(OX)x Polymerization

Reactant
-1 -1 -1
f l u x /xmol MmolL molL /LimolL
- 1 - 1 -4 -1
min. min. χ 10 min.
OH ΔΑ1 log k log k mean
o b b 2 2
Expt. T, C pH Al OH OH A — η C 2+ - f o r ex p t .
Τ T b ~23 Al AlOH
b

A114 25 .0 4 .75
. 0.055 .088 . 107 .96 2..0 1..00 .570 . 107 -4 ,. 50
]\ - 4 ..59
. 093 . 123 1 .01 2..2 .83 .619 . 102 -4 ,.56
A112 25. .0 4 ,. 90 .055 . 115 .216 1 .20 2.,3 .77 .431 . 166 -4 . 19 -4 .. 12
A119 25. .0 5..00 .055 . 136 .281 1 .30 2..0 1..00 .317 .281 -3 .83
\ - 3 ,.81
. 138 .285 1 .34 2., 1 .91 .340 .259 - 3 , .90
A120 25. .0 5..00 . 122 .315 .822 1 .50 2.. 4 .71 .453 .587 -3,
\ "3,.60
.312 . 806 1 . 50 2..3 . 77 .453 .621 -3. .64
A121 25. .0 5..20 . 122 . 366 .871 1 . 58 2.. 4 .71 .209 .618 -3 .31 -3. .25
A113 25. .0 5..20 .055 . 177 .412 1 .35 2.. 4 .71 .140 .293 -3 .46
\ "3 .40
. 176 .409 1 .28 2.. 4 .71 . 125 .290 -3 .41
A117 35. .0 4 .75
. .055 .084 . 188 1 .23 2..0 1..00 .368 . 188 -4 .07 -3. .98
A115 35. ,0 4 ,. 90 .055 . 165 .408 1 .24 2..2 .83 . 190 .339 -3. . 53 -3 .54
A116 10. .0 5..20 .055 . 111 .221 1 . 17 2..2 .83 .431 . 183 -4 . 15 -4 . 17

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
438 CHEMICAL MODELING OF AQUEOUS SYSTEMS II

-3.00

LogC Alb
-4.00
molal

-5.00

-3.00
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

-4.00
LogCAla
molal
-5.00

-6.00
0 200 400 600 800 1,000 1,200 1,400
Time, minutes
Figure 2. Total concentrations of Ala and Alb species measured during titration, experiment
Al 12, pH4.90,T = 25.0°C.

-2.50

- 5.00

Figure 3. Kinetics of formation of Ala at 25.0 °C from pH 4.75 to 5.20. Open circles are mean
values of log ki measured in six experimental runs; vertical lines represent range of log k2 values
computed for runs at pH's indicated. Units for k2 are sec. .

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
33. H E M & ROBERSON Aluminum Hydrolysis Reactions and Products 439

There is also a strong temperature effect on the rate constant. Applying the relevant data in
Table V to the Arrhenius relationship:

(log k - log ki) χ 2.303 R Τι T2


2

where Ea is the energy of activation, k and Τ represent rate constants and kelvin temperatures,
respectively, and R is the gas constant, gives Ea values of 25.6, 27.3 and 21.6 kcal for pH's of 4.75,
4.90 and 5.20, respectively. The thermodynamic significance of these numbers is obscured by the
strong effect of temperature on H activity, and the empirical nature of the rate constant. However, the
low Ea obtained at pH 5.20 may be related to the formation of some Ale at that pH during titration.

DISCUSSION OF KINETIC RESULTS


Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

The effect of using different rates of Al addition can be observed in the data given in Tables IV and V.
It was assumed that the calculated rate constant would not be significantly affected by using different
Al addition rates. This required that the rate chosen be adequate to achieve the necessary reaction
affinity but not so fast that it would prevent a steady-state from being reached. Supersaturation that
may occur in the solution in close proximity to the reagent delivery points proably does not have an
important effect and the measured pH is assumed to represent conditions throughout the solution
during titration. The slow addition rates, the dilute nature of the added solutions and continuous
mechanical stirring all help minimize local supersaturation. In addition, the formation of Ala species
is fast and reversible, while the irreversible steps to form Alb and Ale are probably not fast enough to
progress significantly during the short time any given parcel of the solution is passing through the area
affected by incomplete mixing of reagents.
n
Approximately doubling the rate of Al addition in titration increased the average value of k2
by 0.21 log unit at pH 5.00 and by 0.15 log unit at pH 5.20 (Table V). This suggests that the system
may not have been totally self-compensating for this factor. However, the effect on the rate constants
is probably within uncertainty limits that are imposed by various other factors that affect the
experimental results.
In titrations at pH 5.20 and 5.00 the size of Alb units, as indicated by the OHb/Alb ratio seems
to reach a maximum at 2.4, which could be interpreted as a boundary in the transition of Alb to Ale,
microcrystalline gibbsite. Some additional support for this hypothesis can be obtained from
experiments of Parthasarathy and Buffle (8), who used ultrafiltration to separate polymeric Alb from
Ale. They observed that the polymeric material passed through a filter having a nominal porosity of
20 X. Preliminary experiments we have made with this technique indicated that our Alb fraction also
can pass through this membrane, but Ale docs not.
+
Figure 4 includes a structure comprising four rings each having 6 Al ions connected by (OH>
bridges that would have the formula Α1ΐ6(ΟΗ>8 . This unit has a diameter somewhat less than 20 °A (I),
+

and its OH/ΑΙ ratio is 2.38. It is also of interest to note that such a unit has 10 Al ions in external
sites where an unsatisfied unit of positive valence is available. An additional shared (OH)2 bridge
could be established at such sites as the unit grows laterally. There are geometric constraints on the
ways in which the six-membered ring units can combine. The A1IO(OH)38 unit can be formed from 2
rings and 2 dimeric Ah(OH)2 units as shown in the drawing by adding 4 more (OH> bridges, and
other combinations could be imagined but it cannot form by the interaction of 4 Α1β(ΟΗ)ΐ2 units. The
net charge per Al in the Al i6(OH>8 unit is + 0.63. Parthasarathy and Buffle (8) determined that their
Alb polymer had a net charge of 0.53 or 0.56, depending on the method of determination. They
favored the composition Α1ΐ3θ4(ΟΗ)24 (net charge + 0.54 per Al). As will be noted later, we do not
believe this form of the polymer is compatible with a reaction path leading to crystalline gibbsite.

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
440 CHEMICAL MODELING OF AQUEOUS SYSTEMS II

CHARACTERIZATION OF Alb

The differing chemical behavior of hydroxide bound in monomeric species from that bound in Alb,
the polymeric material, is evidenced in various ways. From the viewpoint of coordination chemistry,
an explanation for this difference is the development of a bridging structure in which hydroxide ions
in pairs link adjacent A l ions. The structure of gibbsite can be represented in three dimensions by a
double layer of close-packed spheres, each representing an OH ion. Smaller spheres representing A Î +

ions occupy octahedral coordination positions in the interstices between layers. Each A l is
coordinated with 6 OH ions, but to maintain a charge balance each A l ion shares the total 6 negative
charge units on these OH ions with 3 other A l ions. The resulting structural pattern can extend
indefinitely in the a and b directions. The individual A l ions occupy only 2/3 of the octahedral sites
and the structural units can be schematically represented as an hexagonal ring with the sides
representing pairs of bridging OH ions and points marking the positions of A l (Fig. 4a). The single
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

six-membered ring would have the formula Ale(OH)i2 with each A l still having a single unsatisfied
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

positive charge.
This structural unit for polymeric material was suggested by Hsu and Bates (14), and was
adopted in our earlier work (12). It also was used by S toi et al. (15) as a basis for explaining
precipitate growth in hydrolyzing A l solutions. It is the basis for the representation of Alb as
Alc(OH)i2 in the equation for conversion of Ala to Alb (equation 2). In effect, the rate constants
determined here represent the rate of converting free or monomeric OH into bridging OH.
Growth mechanisms represented by Figure 4b show how Ak(OH)i2 units might be joined by
establishing new (OH> bridges (represented by dashed lines) between the rings. The gibbsite
structure of Ale entails stacking of the layers in the c direction, following an hexagonal close packing
arrangement, where hydrogen bonding between adjacent double layers aligns them in an ab, ba,
sequence (Fig. 4c, after (16)). Estimates by Smith and Hem (1) of the thickness of Ale particles
suggest they would contain three or more double layers.
The structural model for the principal polynuclear aluminum hydroxide ion given by Baes and
Mesmer (17) is for a species having the composition Α1ΐ3θ4(ΟΗ)24. It involves 3 coalesced Α1ό(ΟΗ)ΐ2
+
rings with an additional A l ion in the center of the structure that is in tetrahedral coordination with 4
oxygen ions that are included in the AU rings. Presumably these originally were OH ions but for
some reason have been deprotonated. This structure was originally proposed by Johansson (18), who
identified units of this type by X-ray diffraction in precipitates of basic aluminum sulfate and selenate
prepared by adding Na2S04 or Na2Se04 solutions to a partly hydrolized AlCh solution (ΟΗτ/ΑΙτ =
2.5). No gibbsite was reported in these precipitates.
The Α1ΐ3θ4(ΟΗ)24 unit would appear unlikely to assimilate readily into a gibbsite structure, as
the latter does not include Ο ions or 4 coordinated A l . Furthermore, OH ions of the complex are
arranged in cubic close packing rather than the hexagonal close-packing of gibbsite. Aluminum
oxyhydroxides might form under some experimental conditions and at higher A l concentrations than
we have used. Parthasarathy and Buffle (8) postulated that the polymeric Al(OH)x species they
27 x
synthesized was predominantly Α1ΐ3θ4(ΟΗ)24. The A l NMR spectra of solutions containing relatively
high concentrations of Alb (0.1 molal in total Al) from which Ale had been removed by ultrafiltration
gave a peak they attributed to tetrahedrally coordinated A104 groups.
Work by Bertsch, Anderson and Layton (19) that included NMR studies indicates that the AI13
-3 0

polymer is formed in solutions having A l concentrations greater than 10 molal or that have high
ΟΗτ/ΑΙτ ratios. These investigators suggested that Al(OH> is a required precursor for the formation
of the Α1ΐ3θ4(ΟΗ)24~ unit.
It seems unlikely that significant amounts of the AI13 polymeric species cited above would
have been produced under our experimental conditions. Perhaps the reluctance of Alb to convert to
Ale during aging of some of the solutions could be attributed to the extensive structural
rearrangements that conversion of the AI13 polymer to gibbsite would entail.

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
33. H E M & ROBERSON Aluminum Hydrolysis Reactions and Products 441

78 a
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

AI (OH), "
6 2
64

Figure 4(a). Schematic representation of Al6(OH)i2 complex. Al ions at corners; connecting lines
represent pairs of bridging O H ions (not to scale).

16a
AI (0H)3
|6 8
,(

(b). Schematic representation of A1IO(OH)38 polymeric ion. Solid lines outline original polymeric
units and dimers; dashed lines are new double O H bridges formed between original units.

(c). Gibbsite double-layer stacking arrangement.

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
442 CHEMICAL MODELING OF AQUEOUS SYSTEMS II

A study of polymerization kinetics and species structure was made by Stol et al. (15) using a
batch titration technique and characterization of polymers by light-scattering. Although their
experimental conditions differed substantially from those used in this study, many of their conclusions
are in general accord with our own. They suggested that above a critical value for the ratio COHb/CAlb
of «2.5 the positive charge on the edges of hexagonal Alb is no longer a strong deterrent to
aggregation of Alb species to form crystalline solids with the gibbsite-type layered structure. By this
interpretation, the development of Ale, or microcrystalline gibbsite does not become a major factor in
the hydrolysis process in titration done between pH 4.75 and 5.20. Values for the OHb/Alb ratio given
in Table V reached a maximum of 2.4.
A later study by DeHek et al. (20) explored the effect of SOÎ ions on the precipitation
mechanism. These investigators suggested that SO4 could form surface groups on the Alb structures
that catalyze the formation of precipitates. The role of SO4 in aluminum hydrolysis in natural systems
also was thought by Nordstrom (4) to be an important one.
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

ENVIORNMENTAL AND GEOCHEMICAL SIGNIFICANCE OF AlOH POLYMERIZATION

Effect of Mineral Surfaces on Al Behavior

It seems likely that there are some natural aqueous systems, especially those impacted by acidic
precipitation or discharges of acid mine drainage, in which polymeric Al species like those studied in
this work will be formed. Most of these systems will contain solid surfaces, and a tendency for
aluminum polymeric species to nucleate and precipitate on certain types of mineral surfaces was
documented by Brown and Hem (21). Conclusions relevant to the probable course of the aluminum
hydroxide polymerization process in natural water may be summarized as follows:
1. The behavior of monomeric (Ala) species is explainable, in general, as a gradual approach
to gibbsite equilibrium except for solutions that contained excessive concentrations of dissolved silica
that had been leached from the mineral substrates. The reaction affinities for gibbsite precipitation are
similar to those observed at similar aging times in the absence of surfaces. In the silica-rich solutions,
precipitation of clay-mineral precursors could have occurred.
2. Rate of the Alb polymer growth to become Ale appears not to be affected significantly by
the presence of mineral surfaces. The same general rate and first-order disappearance of Alb occurred
in both the presence and absence of surfaces, after an initial induction period. However, in the
absence of surfaces the growth process did not attain a steady well-defined rate until after about 3
weeks of aging. In the presence of surfaces there was an almost immediate decrease in Alb that
continued until the end of the 100 day aging period.

Observations in Natural and Anthropogenicaly-Influenced Systems

Removal of color and turbidity caused by particulate matter is commonly accomplished, during the
treatment of water for municipal and industrial purposes, by adding a solution of Al , to the raw water
in a holding tank or mixing facility while maintaining the pH at a level favorable for aluminum
hydrolysis. The particulate material in the raw water is coagulated by the aluminum hydrolysis
species and can be removed by filtration. Much of the literature on aluminum hydrolysis is directly or
indirecUy related to the chemistry involved in this process.
Depending on the various changing properties of the input water, there can be an aluminum
residual probably consisting of polymeric aluminum hydroxide in the treated water that is not
removed by filtration. Miller et al. (22) compared total aluminum concentrations (determined by an A-
A procedure) in treated and untreated water from 184 United States public water supply systems and
found substantially more A l in the treated than in the untreated water in many instances. Similar
conditions can be seen in data collected earlier on the chemical composition of water supplies of the

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
33. H E M & ROBERSON Aluminum Hydrolysis Reactions and Products 443

100 larger U. S. cities (22). The additional aluminum in these treated waters probably was a form of
Alb or Ale, but identification of specific forms was not attempted in these studies. The failure of ion-
exchange column demineralizer units to remove aluminum from tap water furnished by public utilities
(24) does suggest that it is polymeric, and possibly in the colloidal size range. A later survey (25)
provides additional information on possible causes of residual Al in treated water.
A commonly observed effect of acidic precipitation is an increased concentration of aluminum
in water of streams and lakes. Water of lakes that received acidic runoff may increase in transparency
(26) owing to coagulation and settling of suspended organic and inorganic particulates. This effect is
probably related to aluminum hydrolysis, as it is the same as the coagulation step in water-supply
treatment cited earlier.
A study of the chemistry of a small California stream that receives acid mine water (27)
showed that above pH 4.9 the dissolved aluminum concentration appeared to be controlled by
aluminum hydroxide solubility equilibria, but below pH 4.6 aluminum hydroxy-sulfates apparently
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

predominated. This study also cited data for acid mine drainage in the Appalachian region and lakes
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

affected by acid precipitation that showed a similar pattern of aluminum behavior.


A procedure developed by Barnes (28), and later modified by Driscoll (29), has been used to
separate monomeric reactive species of A l from polymeric and colloidal forms. In tests of the
procedure, (28) extractable A l concentrations in a series of natural-water samples that had been
filtered through 0.10 μιη pore-size membranes, acidified to pH 2.0 and stored for several weeks, were
compared with extractable Al in aliquots of the same filtered samples that were not acidified and on
which the extraction was done immediately after filtration. The results indicated that acidification and
storage increased the extractable Al concentration by nearly a factor of 2 in these samples.
All the samples included in the Barnes (28) study were from groundwater sources and most
had pH's near neutrality. The acid soluble fraction passing through the 0.1 μηι pores of the filter
probably was microcrystalline gibbsite and polymeric aluminum hydroxide macro ions. The kinetics
of dissolution of polymeric Al(OH)x species in acid were studied by Hem (30). Material passing
through 0.10 μηι porosity filters showed psuedo first-order behavior when dissolution pH was held
constant, and the half-time for dissolution was near 10 hours at pH 3.0 and about 12 minutes of pH
2.0. The polymeric species studied in the kinetic experiments were prepared by a batch mixing
technique similar to that used by Smith and Hem (1). The solutions were aged before filtration for
periods of a few days up to two weeks. Probably a larger fraction of the aluminum present in surface
water could have characteristics of Alb than is to be expected in ground water. Currently, available
data are not sufficient to make any quantitative statement. Campbell et al. (31), developed procedures
for aluminum speciation that utilize filtration and chelating resin treatment of surface water samples
and suggested that aluminum hydroxide polynuclear cations less than 60 °A in diameter may be taken
up by the resin along with monomeric complex species.
Kennedy et al. (32). reported the effect of using different filtration and preservation techniques
on the aluminum concentration determined in the filtrate of a sample collected at high flow from the
Mattole River of northern coastal California. The aluminum concentration determined in the first 2
liters of water filtered through a 0.45 μιη porosity filter and subsequently acidified with ultrapure
nitric acid (final pH 1.5 to 1.7) and stored for 90 days before analysis was more than 2.2 μ π ι ο ^ . An
aliquot of the same original water filtered through a 0.10 μηι porosity filter and given the same
subsequent treatment yielded an aluminum concentration somewhat less than 0.6 μηιο^Ι. This
concentration also was measured within an analytical uncertainty of about ± 0 . 2 μιηο^Ι in a portion of
the 0.10 μηι filtrate that was not acidified but from which Al was extracted by an organic reagent (8-
quinolinol dissolved in methyl isobutyl ketone) immediately after filtration. In a companion study,
Jones et al. (33) calculated that their Mattole River water sample was more than an order of
magnitude supersaturated with respect to kaolinite, which should have been the most stable aluminum
bearing solid. An electron micrograph of particulate matter from a Mattole River sample that had

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
444 CHEMICAL MODELING OF AQUEOUS SYSTEMS II

passed through a 0.45 μπι filter was published by Hem et al. (34). Hexagonal units resembling
gibbsite are visible in this micrograph, but their composition was not determined.
Whether the Mattole River sample of Kennedy et al. (32) contained a significant amount of Alb
cannot be determined from the data they presented. Colloidal particulates that perhaps contained Ale
were probably present in material passing through the 0.45 μπι filter. The extraction procedure used
in their experiments involved a 15 minute contact time of the 8 quinolinol reagent with the sample
making it likely that some, at least, of the Alb fraction would be extracted. The method of Barnes (28)
uses the same extractant but only a 10 sec. contact time, in order to minimize extraction of polymeric
Al species.
Effects of aluminum in low pH solutions on aquatic biota, especially fish, and on land plants
growing in acidic soil have been studied by various investigators (35-38). Although much of the work
has assigned the observed toxicity to monomeric dissolved forms of Al, the behavior of polymeric
species in biochemical systems is probably in need of more careful study than it has so far received.
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

Schindler Q9) in a review of the topic of environmental impacts of acid rain noted that aluminum is
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

highly toxic to fish and quoted literature references to the effect that polymeric and colloidal
aluminum hydroxide species may physically obstruct gill membranes and cause asphyxiation. The
maximal pH range for this effect was reported as 5.2 to 5.4.

Implications of Polymerization Mechanisms and Kinetics for Occurrence of Alb in Streams

The experimental data given and reviewed here show that there are critically important changes in the
hydrolysis behavior of aluminum between pH 4.5 and 5.5. Early stages of polymerization and
precipitation of aluminum hydroxide and related secondary minerals occur near pH 5.0 and can be
greatly accelerated or retarded by small pH changes.
It would appear likely that small amounts of Ale could be present as colloidal solid particles in
many river waters. Conditions favorable for stabilizing polymeric Alb ionic species are readily
attainable in the laboratory, as our study has demonstrated. The maximum Alb yields were in
solutions titrated at pH 5.00, and total monomeric aluminum concentrations were held near 10
molal. This high an Ala concentration in natural water could be reached through dissolution of soil
minerals at a pH near or below 4.50.
The requirement that the solution be supersaturated with respect to gibbsite to form Alb (A23 >
1.0 kcal) would be difficult to meet at pH's much below 4.75, owing to the rapid increase in
equilibrium solubility of Al and decrease in proportion of the AlOH species as pH decreases. For
significant amounts of unreacted polymeric Alb to be present in surface water requires that there be a
prior history of low-pH inflow and a supply of dissolved Al, along with low levels of aluminum-
complexing species other than OH. Once formed, the smaller polymeric units might well persist for a
considerable time. The kinetics of Alb formation in the open-system laboratory experiments are
directly relevant toward improving our understanding of the environmental behavior of this element.

LITERATURE CITED

1. Smith, R. W.; Hem, J. D. U.S. Geol. Surv. Water-Supply Pap. 1827-D, 1972, D1-D51.
2. Hem, J. D. U.S. Geol. Surv. Water-Supply Pap. 1827B, 1968, B1-33. p.
3. Lind, C. J.; Hem, J. D. U.S. Geol. Surv. Water-Supply Pap., 1827G, 1975, G1-G83.
4. Nordstrom, D. K. Geochim et Cosmochim Acta, 1982, 46, 681-692.
5. Smith. R. W. In Nonequilibrium, Systems in Natural Water Chemistry; Hem, J. D., Ed.;
Amer. Chem. Soc., Washington, D.C, 1971, 250-279
6. Okura, T.; Goto, K.; Yotuyanagi, T. Analytical Chem., 1962, 34, 581-582.
7. Turner, R. C. Canadian J. of Chem., 1969, 47, 2521-2527.
8. Parthasarathy, N.; Buffle, J. Water Research, 1985, 19, No. 1, 25-36.

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
33. HEM & ROBERSON Aluminum Hydrolysis Reactions and Products 445

9. Bersillon, J. L.; Hsu, P. H.; Fiessinger, F. Soil Sc. Soc. Amer. J., 1980, 44, 630-634.
10. Jardine, P. M.; Zelazny, L. W. Soil Sci. Soc. Amer. J., 1986, 50, 895-900.
11. Seip, H. M.; Muller, L.; Naas, A. Water. Air and Soil Pollution, 1984, 23, 81-95.
12. Hem, J. D.; Roberson, C. E. U.S. Geol. Surv. Water-Supply Pap., 1827A, 1967, A1-A55.
13. Parks, G. A. Amer. Mineralogist, 1972, 57, 1163-1189.
14. Hsu, P. H.; Bates, T. F. Mineralogical Magazine, 1964,33,749-768.
15. Stol, R. J.; Van Helden, A. K.; De Bruyn P. L. J. of Colloid and Interface Sci., 1976,
57, 115-131.
16. Wefers, K.; Misra, C. Alcoa Tech. Pap. No.19,revised, 1987, 1-92.
17. Baes, C. F.; Mesmer, R. E. Wiley-Interscience: New York, 1976, 489.
18. Johansson, G. Svensk Kemisk Tidskrift, 1963, 2, 6-7.
19. Bertsch, P. M.; Anderson, Μ. Α.; Layton, W. J. Preprint. Env. Chem. Div. 194 Amer. Chem.
Soc. Mtg., 1987, 142-143.
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org

20. DeHek, H.; Stol, R. J.; Debruyn, P. L. J. Colloid Interface Sci., 1978, 64, 72-89.
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

21. Brown, D. W.; Hem, J. D. U.S. Geol. Surv. Water-Supply Pap. 1827-F, 1975, F1-F48.
22. Miller, R. G.; Kopfler, F. C.; Kelty, K. C.; Stober, J. Α.; Ulmer, N. S. Amer. Water Works
Asso. J., 1984, 76, 84-91.
23. Durfor, C. N.; Becker, E. U.S. Geol. Surv. Water-Supply Pap. 1812, 1964, 364 p.
24. Kerr, D. N. S.; Ward, M. K.; Arze, R. S.; Ramos, J. M.; Grekas, D.; Parkinson, I. S.; Ellis, H.
Α.; Owen, J. P.; Simpson, W.; Dewar, J.; Martin, Α.; McHugh, M. F. Kidney International,
1986, 29, Supp. 18, S-58-S-64.
25. Letterman, R. D.; Driscoll, C. T. Amer. Water Works Asso. J., 1988, 80, No. 4, 154-158.
26. Effler, S. W.; Schofran, G. C.; Driscoll, C. T. Canadian J. of Fisheries and Aquatic Sci., 1985,
42, 1707-1711.
27. Nordstrom, D. K.; Ball, J. W. Science, 1986, 232, 54-56.
28. Barnes. R. B. Chem. Geol1975,15,177-191.
29. Driscoll, C. T. Inter. Nat. J. of Enviro. Analytical Chem. 1984, 16, 267-283.
30. Hem, J. D. In Leaching and Diffusion in Rocks and Their Weather-Products; SS. Augustithis,
Ed.; Theophrastus Publications SA, Athens, Greece, 1983, 51-62.
31. Campbell, P. G. C.; Bisson, M.; Bougie, R.; Tessier, Α.; Villenuve, J. P. Analytical Chem.,
1983, 55, 2246-2252.
32. Kennedy, V. C.; Zeleger, G. W.; Jones, B. F. Water Resources Research, 1974,10,785-790.
33. Jones, B. F.; Kennedy, V. C.; Zellweger, G. W. Water Resources Research, 1974, 10, 791-793.
34. Hem, J. D.; Roberson, C. E.; Lind, C. J.; Polzer, W. L. U.S. Geol. Surv. Water-Supply Pap.
18227-E, 1973, E54.
35. Baker, J. K. In Acid Rain/Fisheries; Johnson, Ε. E., Ed.; Amer. Fisheries Soc: Bethesda,
Maryland, 1982; p. 165.
36. Driscoll, C. T.; Baker, J. P.; Bisogni, J. J.; Schofield, C. L. Nature. 1980, 284, 161-164.
37. Foy, C. D. In The Plant Root and its Environment; Carson, E. W., Ed.; University of Virginia
Press: Charlottesville, Virginia, 1974, p 601.
38. Parker, D. R.; Zelazny, L. W.; Kinraide, T. B. Proc. 194 Amer. Chem. Soc. Mtg., 1987,
p. 369-372.
39. Schindler, D. W. Science, 1988, 239, 149-157.
40. Hemingway, B. S.; Robie, R. A. Geochim. Cosmochim. Acta, 1977, 41, 1402-1404.
41. Raupach, M. Australian J. of Soil Res., 1963, 1, 28-35.
42. Hemingway, B. S.; Haas, J. L. Jr.; Robinson, G. R. Jr. U.S. Geol. Surv. Bull. 1544, 1982,
70 pp.
43. Hemingway, B. S.; Robie, R. Α.; Kittrick, J. A. Geochim. Cosmochim. Acta, 1978, 42,
1533-1543.
44. Wagman, D. D.; Evans, W. H.; Parker, V. B.; Halow, I.; Bailey, S. M.; Schumm, R. H.
Nat. Bureau of Standards Tech. Note 270-3, 13, 207-208.

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.
446 CHEMICAL MODELING OF AQUEOUS SYSTEMS II

45. May, H. M.; Helmke, P. Α.; Jackson, M. L. Geochim. Cosmochim. Acta. 1979, 43, 861-868.
46. Singh, S. S. Proc. Soil Sci. Soc. of Amer., 1974, 38, 415-417.
47. Kittrick, J. A. Proc. Soil Sci. Soc. of Amer., 1966, 30, 595-598.
48. Nordstrom, D. K.; Plummer, L. N.; Langmuir, D.; Busenberg, E.; May, H. M.; Jones, B. F.;
Parkhurst, D. L. This volume, 1989.
49. Ackerman, T. Zeitschrif fur Elektrochemie 1958, 62, 411-419
RECEIVED August 31, 1989
Downloaded by CORNELL UNIV on August 18, 2016 | http://pubs.acs.org
Publication Date: December 7, 1990 | doi: 10.1021/bk-1990-0416.ch033

Melchior and Bassett; Chemical Modeling of Aqueous Systems II


ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Das könnte Ihnen auch gefallen