You are on page 1of 15

Chemical Geology 165 Ž2000.

267–281
www.elsevier.comrlocaterchemgeo

Dissolution of forsteritic olivine at 658C and 2 - pH - 5


)
Yang Chen, Susan L. Brantley
Department of Geosciences, PennsylÕania State UniÕersity, UniÕersity Park, PA 16802, USA
Received 29 July 1998; accepted 12 August 1999

Abstract

Dissolution experiments with forsteritic olivine ŽFo 91 . were conducted in a batch reactor at a temperature of 658C
between pH 2 and 5. Constant pH was maintained by using a pH-stat technique. The following dissolution rate law for
forsteritic olivine at 658C and acid pH was derived based on the experimental results normalized to the initial surface area:

0.70
r s 10y8 .51 Ž a Hq .

where r is rate in mol forsteritic olivinercm2rs. Our results, combined with data from the literature for forsteritic olivine
dissolution at 258C, show that the pH-dependence of forsteritic olivine dissolution is temperature-dependent. As temperature
increases, the dissolution rate of forsteritic olivine becomes more pH-dependent, which is consistent with a surface
protonation model for dissolution at pH - pH pznpc . The activation energy of dissolution, Ea , has been estimated based on our
results and literature data at 30 " 4 kcalrmol. We also observed that at pH between 2 and 4 and at 658C, the release rates of
Mg, Si and Fe were stoichiometric. However, at pH 5, the release rate of Fe was slower than that of Mg and Si, probably
due to oxidation of the mineral surface. Results of several dissolution experiments in the presence of Al in solution ŽpH 3,
0.03–10 ppm Al. show that the dissolution rate of forsteritic olivine under the experimental conditions is independent of Al
concentration within the experimental error. This result may indicate that Al does not form strong crosslinks to the
unpolymerized surface of orthosilicates such as forsteritic olivine, and therefore does not retard the dissolution of this phase
under the experimental conditions. q 2000 Elsevier Science B.V. All rights reserved.

Keywords: Forsteritic olivine; Dissolution kinetics; Simple oxides

1. Introduction the reactive solution species are Hq, OHy or other


suitable ligands Že.g., Stumm and Furrer, 1987.. The
Extensive research on dissolution kinetics shows
surface protonation behavior of simple oxides such
that for simple oxides, dissolution is controlled by
as Al 2 O 3 , TiO 2 and Fe 2 O 3 as a function of pH and
surface reaction at the oxide–water interface, where
temperature have been constrained Že.g., Berube and
DeBruyn, 1968; Furrer and Stumm, 1983; Carroll-
)
Corresponding author. Tel.: q1-814-863-1739; fax: q1-814- Webb and Walther, 1988; Machesky, 1989; Brady
865-3191; e-mail: brantley@geosc.psu.edu and Walther, 1992.. Dissolution behavior as a func-

0009-2541r00r$ - see front matter q 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 0 0 9 - 2 5 4 1 Ž 9 9 . 0 0 1 7 7 - 1
268 Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281

tion of pH is less well constrained for mixed oxides Polymerization of the dissolving silicate phase
such as ortho-, ino-, phyllo-, and tectosilicates. may also be important with respect to inhibition by
Several authors have argued that mineral dissolu- aluminum. Several workers have shown ŽChou and
tion will be pH-dependent for pH - pH pznpc or pH ) Wollast, 1985; Casey et al., 1988; Oelkers et al.,
pH pznpc Že.g., Furrer and Stumm, 1986.. Here, 1994; Chen and Brantley, 1997. that Al in solution
pH pznpc stands for pH of point of zero net proton affects dissolution of feldspar, perhaps because it
charge, or the pH at which the concentration of repolymerizes the silica-rich feldspar surface. Fewer
positively charged surface sites equals the concentra- workers have investigated Al inhibition on less poly-
tion of negatively charged surface sites of the min- merized silicate phases. One of the purposes of this
eral in the absence of specifically adsorbed cations study was to look at the pH-dependence of a simple
or anions. In general, the pH dependence of silicates orthosilicate by conducting an experimental study of
has been described by fitting the rate of dissolution forsteritic olivine dissolution at an elevated tempera-
for pH - pH pznpc to a simple rate equation for rate ture, and to compare the results with the literature
data at constant temperature: data for dissolution of this phase at room tempera-
ture. The second purpose was to test whether dis-
solved Al inhibits olivine dissolution.
log r s log k y n pH Ž 1.

where r is the dissolution rate, k is the apparent rate


constant and n is the apparent reaction order with 2. Dissolution of forsteritic olivine
respect to activity of aqueous Hq.
Brady and Walther Ž1992. and Casey and Sposito The dissolution kinetics of forsteritic olivine at
Ž1992. suggested, based on thermodynamic argu- 258C have been studied by many workers Že.g., Luce
ments, that the pH-dependence of mineral dissolution et al., 1972; Sanemasa et al., 1972; Siever and
should increase with temperature at pH - pH pznpc . Woodford, 1979; Grandstaff, 1980, 1986; Siegel and
However, the effects of temperature on the pH-de- Pfannkuch, 1984; Blum and Lasaga, 1988; van Herk
pendence of silicate dissolution are not well-estab- et al., 1989; Sverdrup, 1990; Wogelius and Walther,
lished. For example, Hellmann Ž1994. published data 1991; Varadachran et al., 1994; Jonckbloedt, 1997;
suggesting that n increased with increasing tempera- Rosso and Rimstidt, in press.. Luce et al. Ž1972.
ture Žfrom 100 to 3008C. for albite; however, reanal- reported no pH dependence for olivine dissolution,
ysis of his data with our data suggested to us that n but his experimental protocol left ultrafine particles
may remain constant from 58 to 3008C for dissolu- in the starting materials and, as a result, all dissolu-
tion of this phase ŽChen and Brantley, 1997.. In tion was parabolic. Grandstaff Ž1980; 1986. dis-
contrast, our experiments at pH - pH pznpc show that solved forsteritic olivine ŽFo 83 . in HCl q KCl at pH
both diopside and anthophyllite dissolution become 3 to 5 between 18 and 498C. His results showed that,
more pH-dependent at 908C than at 258C ŽChen and at 258C, dissolution rate decreases from 10y1 4 to
Brantley, 1998.. The increase in pH-dependence with 10y1 6 molrcm2rs as pH increases from 3 to 5,
temperature is greater for dissolution of diopside indicating a first order dependence of rate on  Hq4 .
than anthophyllite over the same pH range ŽChen He also determined that the activation energy of
and Brantley, 1998.. If the effect of temperature on dissolution based on regression of log rate Žat pH
pH-dependence is related to the degree of poly- 3.25. vs. 1rT equals 9.1 kcalrmol. However, the
merization of the silicate, we might expect to see an rates reported by Grandstaff Ž1980; 1986. are 1–2
even stronger temperature effect on the dissolution orders of magnitude lower than the rates reported
of silicates that are even less polymerized, such as later for forsteritic olivine and other orthosilicates at
forsteritic olivine, at pH - pH pznpc . For example, similar temperature and pH conditions Že.g., Blum
Westrich et al. Ž1993. report a strong change in pH and Lasaga, 1988; Sverdrup, 1990; Wogelius and
dependence with temperature for several orthosili- Walther, 1991; Westrich et al., 1993.. Some of this
cates over the range T s 25 to 658C. discrepancy may have been related to an erroneously
Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281 269

measured BET surface area Žas suggested by Mur- reported for the mineral composition. According to
phy, 1985; Wogelius and Walther, 1991, 1992.. Wogelius Žpers. comm.., the nonstoichiometry of the
Blum and Lasaga Ž1988. also reported a linear product solution at the end of the experiment might
relationship between log dissolution rate and pH for have been caused by precipitation.
dissolution of forsteritic olivine ŽFo 93 . Ž n s 0.56.. In addition to surface protonation, other factors
Their results show that at 258C, when pH increases such as Al inhibition may also play an important role
from 2 to 4, dissolution rate decreases from 10y1 2 to in controlling the dissolution of some aluminosili-
10y1 3 molrcm2rs. Forsteritic olivine dissolution ex- cates ŽChou and Wollast, 1985; Oelkers et al., 1994;
periments at 258C were also conducted by Sverdrup Chen and Brantley, 1997.. However, it is unclear
Ž1990.. However, little information is given about whether Al inhibition affects all silicate dissolution.
the experiments and materials, and rates presented by According to Casey et al. Ž1988., Al inhibition of
this author are scattered. feldspar dissolution is the result of polymerization in
Results similar to those of Blum and Lasaga which adsorbed Al 3q retards the dissolution by
Ž1988. were obtained for forsteritic olivine ŽFo 91 . adding crosslinks to the silica network. No investiga-
dissolution at 258C by Wogelius and Walther Ž1991.. tions of the effect of dissolved Al on olivine dissolu-
These workers measured rates of dissolution in ex- tion have been reported; therefore, it is not clear
cellent agreement with the results of Blum and Lasaga whether Al 3q in solution would also crosslink the
Ž1988.. Using the results reported by both groups unpolymerized surface and retard the dissolution of
Žwithout the data measured at pH 5.7., forsteritic this phase.
olivine dissolution at 258C and pH below neutral can
be expressed using the equation:
3. Methods
y11.01 0.49
r s 10 Ž aHq . Ž 2.
Forsteritic olivine ŽFo 91 ., from San Carlos, Ari-
zona, was purchased from Ward’s Natural Science
where r is the rate, expressed in mol forsteritic
Establishment ŽTable 1.. The starting material Žap-
olivinercm2rs.
proximately 5 = 5 = 5 mm., was hammer-crushed
Westrich et al. Ž1993. show some dissolution data
and ground using an agate mortar to a fine powder,
for forsteritic olivine Žbut without reporting rate val-
and then dry-sieved to obtain a size fraction of
ues. and report a value of n based upon their own
100–200 mesh Ž150 to 75 mm.. The powder was
data and that of Wogelius and Walther Ž1991. of
then ultrasonically cleaned in ultrapure acetone, dried
0.54. Westrich et al. Ž1993. report that n varies from
in an oven at 1058C, and stored in a dessicator before
0.4 to 0.6 for six different compositions of orthosili-
experiments. Surface area of the powder samples
cate minerals.
before and after experiments were measured using
Wogelius and Walther Ž1992. also conducted
3-point BET method with Kr as the adsorbent
forsteritic olivine dissolution experiments at 658C.
ŽQuantachrome Quantasorb Model OS-10.. Observa-
Two experiments at pH below neutral were run at
tions by Wogelius and Walther Ž1992. and in our
658C ŽpH 1.8 and 6 for up to 25 h.. Their results
laboratory revealed no evidence for impurity phases.
show that between 258 and 658C the pH-dependence
of the rate does not change significantly, and they
estimated the apparent activation energy for dissolu-
tion of forsteritic olivine Žbased on a plot of log rate
Table 1
vs. 1rT . to be 19 " 2.5 kcalrmol. ŽBecause the Sample Composition of San Carlos Olivine
pH-dependence did not change, this Ea is not pH-de-
MgO SiO 2 FeO NiO Total MgrSi FerSi
pendent.. However, dissolution at pH 6 was incon- Žwt.%. Žwt.%. Žwt.%. Žwt.%. Žmole ratio. Žmole ratio.
gruent. The measured ratio of MgrSi in solution at
49.24 40.72 9.21 0.39 99.54 1.8 0.19
the end of their pH 6.0 experiment is 11.3, which is
U
much higher than the stoichiometric ratio of 1.8 Data from Wogelius and Walther Ž1992..
270 Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281

The reactor used in this study, a 2-l batch reactor probe was inserted into the solution to monitor the
made of polyethylene, is similar in design to those change of pH resulting from dissolution Žwith an
used by Casey et al. Ž1988.. The lid of the reactor Orion Model 230A pH meter.. The electrode was
was made of a rubber stopper with holes for installa- periodically calibrated against standard pH solutions
tion of the electrode, the tubing, the rod connector to at 258C. No noticeable drift of the pH electrode in
the stirring bar, and a sampling vent ŽFig. 1.. The measuring buffers was observed for any experiments.
reactor was filled with 2 l of HCl q H 2 O solution The accuracy of pH measurements at 658C was also
adjusted to run pH. One to six grams of powdered checked before experiments. The results of the tests
forsteritic olivine were then added to the reactor, showed that the use of the temperature compensation
which was immersed in a water bath to maintain a probe associated with the pH meter effectively al-
constant temperature of 65 " 18C. The solution in lowed measurement of the correct pH value at 658C.
the reactor was continuously stirred using a magnetic Some previous workers used buffer solutions or
stir bar at either 200 or 350 rpm. To enhance the organic acids to maintain a constant pH during
difference in stirring, a 3 cm long cylindrical stir bar forsteritic olivine dissolution Že.g., Wogelius and
was used for the first case, while a 5 cm cross stir Walther, 1991.. However, buffers may promote or
bar was used for the second case. An ATI-Orion pH inhibit mineral dissolution Že.g., Dove and Crerar,
electrode combined with a temperature compensation 1990; Wogelius and Walther, 1991, 1992; Stillings

Fig. 1. Schematic presentation of the olivine dissolution experiments.


Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281 271

Table 2
Experimental conditions
na s none added.
RUN Temperature Ž8C. pH Ž658C. Al concentration Žmmolrl. Mass of forsterite Žg. Stirring rate Duration Žh.
OV-2 65 4.00 na 4.00 200 5
OV-3 65 2.95 na 2.00 200 5
OV-4 65 2.00 na 1.00 200 2
OV-5 65 5.00 na 6.00 200 5
OV-6 65 2.95 0.37 Ž10 ppm. 2.00 200 5
OV-7 65 2.95 0.1 Ž2.7 ppm. 2.00 200 5
OV-8 65 2.95 0.01 Ž.027 ppm. 2.00 200 5
OV-9 65 2.95 na 2.00 350 5

and Brantley, 1995.. In this study, we therefore used volume of solution that was extracted for chemical
pure HCl q H 2 O. To maintain a constant value of analysis during the experiment was therefore rela-
pH during each experiment, a concentrated HCl solu- tively small Ž- 4%. compared to the initial volume
tion was continuously pumped into the reactor by the of the solution Ž2 l.. Also, removal of liquid was
operation of a peristaltic pump. The rate of HCl partially compensated by the addition of HCl solu-
addition was manually adjusted depending on the tion used to maintain a constant pH Žapproximately
rate of Hq consumption. The pH values of the 20–60 ml of solution was added over the entire
solution were maintained within "0.02 of run pH experiment.. Because of the minimal effect of aliquot
during the experiments. For experiments with added removal, no attempt was made to adjust for the
Al, an aluminum reference solution of 1000 ppm Al change in solution volume. Also, for each experi-
ŽAl dissolved in HCl solution, Fisher Scientific. was ment, approximately 2% of the forsteritic olivine
used to make solutions. powder was dissolved. The change in sample weight
Solution samples were taken from the reactor at was therefore ignored in estimating the dissolution
different intervals. Approximately 6 ml of sample rate. Finally, a small amount of water was lost by
was taken each time, and the maximum number of evaporation from the flask. However, the water loss
samples was 11 for any individual run. The total by evaporation was very small Žless than 2 ml out of

Table 3
Results of dissolution experiments
RUN pH a DG Žkcalrmol. Final areab Žcm2rg. Dissolution rate c Žmol forsteritic olivinercm2rs.
Based on Mg Based on Si
OV-2 4.00 y32.4 560 2.56 " 0.31 = 10y1 2 2.50 " 0.31 = 10y12
OV-3 2.95 y36.9 510 1.75 " 0.22 = 10y1 1 1.74 " 0.21 = 10y11
OV-4 2.00 y43.2 450 8.17 " 1.0 = 10y1 1 8.09 " 0.99 = 10y11
d d
OV-5 5.00 y27.3 570 1.1 = 10y1 2 –3.8 = 10y13 1.2 = 10y12 –5.6 = 10y13
y1 1 y11
OV-6 2.95 y36.6 660 1.52 " 0.19 = 10 1.47 " 0.18 = 10
OV-7 2.95 y36.8 700 1.39 " 0.17 = 10y1 1 1.42 " 0.17 = 10y11
OV-8 2.95 y36.7 630 1.57 " 0.19 = 10y1 1 1.53 " 0.19 = 10y11
OV-9 2.95 nc 540 1.95 " 0.23 = 10y1 1 2.08 " 0.25 = 10y11
a
The listed values are measured pH.
b
Initial specific surface area s 340 cm2rg.
c
Dissolution rate of forsteritic olivine, as mol Mg 1.80 Fe 0.20 SiO4 rcm2rs, normalized to final surface area.
d
See text and Fig. 4 for explanation.
272 Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281

2000 ml, or - 0.1% over a period of 5 h at 658C., dissolution Žthe pH values vary from 2.00 to 5.00..
and therefore is ignored. The extracted solution was Experiments OV-6, OV-7 and OV-8 were used to
immediately filtered through a PTFE filter with pore study the effect of Al on the dissolution of forsteritic
size of 0.45 mm and acidified. Si, Mg, Fe and Al olivine. The Al concentrations ranged from 0.27 to
were analyzed with a PC3000 Leeman Labs ICP 10 ppm Žor 0.01 to 0.37 mmolrl.. One to six grams
atomic emission spectrometer. of forsteritic olivine powder were added to the reac-
tor, depending on the solution pH. Except for the
experiment at pH 2 Žrun for 2 h., all experiments
4. Experimental conditions were run for 5 h ŽTable 2.. Experiments were run for
durations of 2–5 h, based upon the linear Mg and Si
All experiments were conducted at 658C in acid release data reported throughout dissolution of olivine
solutions. Four experiments, OV-2, OV-3, OV-4, by Wogelius and Walther Ž1992. for up to 25 h at
and OV-5, were used to study the pH-dependence of 658C. All experiments were run at a stirring rate of

Fig. 2. ŽA. Changes in concentration of Mg and Si during dissolution of forsteritic olivine at varying pH. ŽB. Changes in concentration of
Mg and Si at pH 2.95, with varying Al concentrations in solution.
Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281 273

Fig. 2 Žcontinued..

200 rpm, except for experiment OV65-9, which was where c Žin molrl. is the concentration, t is time Žin
run at 350 rpm. sec., V is the volume of solution in the reactor Ž2 l.,
A s Žin cm2rg. is the final specific surface area, g Žin
gram. is the mass of forsteritic olivine used, and Õ is
5. Results the stoichiometric coefficient.
The term d crdt in Eq. Ž3. is the slope of the plot
Concentrations of Al, Fe, Mg and Si in solutions
of concentration vs. time, as depicted in Fig. 2 for
of all experiments are summarized by Chen Ž1997..
Mg and Si. Notice that, except for the experiment at
Table 3 lists the surface areas of forsteritic olivine
pH 5 ŽRun OV-5., all experiments show a very good
powder samples before and after experiments and the
linear relation between concentration and time.
dissolution rate, r, calculated based on the following
Therefore, slopes from Fig. 2 were used to calculate
equation:
the dissolution rates ŽTable 3.. These rates vary from
r s Ž d crdt . Vr Ž A s gÕ . Ž 3. 2.5 = 10 y 12 to 8.1 = 10 y 11 mol forsteritic
274 Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281

olivinercm2rs. Good agreement was observed for solution volume change Ž"5%., and surface area
the rates measured at pH 2.9 at two stirring speeds. measurement Ž"10%..
For experiment OV-5 at pH 5, dissolution rate
changed over time. For this experiment, two regres- 5.1. Dissolution stoichiometry
sions were made for each element released, one for
To investigate the stoichiometry of forsteritic
initial dissolution from 20 to 110 min, and the other
olivine dissolution, Mg, Fe, and Si concentrations
for late dissolution between 245 to 300 min. These
were monitored over time and the ratios of MgrSi
rates are 1.12 = 10y1 2 and 3.77 = 10y1 3 mol
and FerSi are plotted in Fig. 3. Results show that, at
forsteritic olivinercm2rs based on Mg release, re-
all experimental conditions, the ratio of MgrSi in
spectively. Table 3 also summarizes the values of the
solution was constant and equal to 1.9 " 0.1, indicat-
DGreaction for all the experiments run at low stirring
ing the congruent dissolution of Mg and Si Žstoichio-
rate Žy43.2 to y27.3 kcalrmol., as calculated by
metric MgrSis 1.8.. However, the release of Fe to
SOLMINEQ ŽKharaka et al., 1988. for the following
the solution is pH-dependent. For pH below 4, we
reaction at 658C:
observed that release of Fe is stoichiometric. At pH
Mg 2 SiO4 q 4Hqs 2Mg 2qq H 4 SiO40 Ž 4. greater than 4, the dissolution rate of forsteritic
olivine estimated based on Fe release is smaller than
An error of about "12% is estimated for the calcu- those based on Mg and Si release, perhaps due to the
lated dissolution rates. The error is estimated based precipitation of an iron-containing phase at higher
on errors in rate of change in solution chemistry pH. At pH s 4, Fe release is retarded for the first
Žincluding chemical analysis and regression, "5%., third of the experiment.

Fig. 3. MgrSi and FerSi release rate ratios for experiments under conditions as indicated. The MgrSi ratio of the forsteritic olivine was
1.8; FerSi was 0.19. All rates were normalized by final surface areas.
Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281 275

acid pH can be represented using the following


equation, using the final surface area:
0.75
r s 10y8.55 Ž aHq . Ž 5.
The log of the rate constant Žs log k s y8.51 "
0.16, using one standard error in the regression of
log rate vs. pH. calculated using the initial surface
area does not differ significantly from the log k
when calculating the rate using the final surface area
Žlog k s y8.55 " 0.12.. Similarly, the value of n
calculated from initial surface area normalization
Ž0.70 " 0.05. does not differ significantly from that
calculated from using final surface area Ž0.75 " 0.04..
Fig. 4. Dependence of forsterite dissolution at pH 2.95 at 658C on We present the rate constants normalized to both
dissolved Al concentration. All rates were normalized by final initial and final surface areas because most experi-
surface areas. mental rates are normalized by initial surface area,
but field weathering rates must be normalized by
final surface areas.
To study the temperature dependence of dissolu-
5.2. Aluminum effect tion rate of forsteritic olivine, the results from this
study are plotted on Fig. 6 Žnormalized by initial
Dissolution rates measured at pH 2.95 are plotted surface area as noted. together with data reported by
against the Al concentrations in Fig. 4. Under the Blum and Lasaga Ž1988. and Wogelius and Walther
given temperature and pH conditions, and for the Ž1991. for dissolution of the same mineral run at
range in concentration of Al Ž0–10 ppm or 0.37 258C. These last two sets of data represent rates
mmolrl., the dissolution rate of forsteritic olivine is normalized to initial surface area. Including errors in
independent of Al. rate measurement Ž"12%., our log rate at pH 2

6. Discussion

6.1. pH-dependence of forsteritic oliÕine dissolution


at 658C

The rates are plotted in Fig. 5 as log rate Žnormal-


ized by final surface area. vs. pH for experiments
performed in the absence of Al. As noted earlier, the
dissolution at pH 5 is non-linear. Therefore, two
dissolution rates are calculated for early and late
dissolution which represent the upper and lower
limits of dissolution rate for the time period between
1200 to 18,000 s at pH 5. These two rates are plotted
using open diamonds in order to distinguish them
from other rate measurements which represented lin-
ear release over the entire run duration.
Fig. 5. pH-dependence of olivine dissolution at 658C. Data plotted
Fig. 5 shows that log dissolution rate is a linear represents rates normalized to final surface area. Line represents
function of pH over the pH range of 2 to 4. The linear regression of all data - pH 5. Rate is expressed as mol
dissolution rate law for forsteritic olivine at 658C and forsteritercm2 s.
276 Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281

pendence of dissolution rate with temperature Žsee


Chen and Brantley, 1998. and is similar to the effect
of temperature as observed on dissolution of three
other orthosilicate minerals ŽWestrich et al., 1993..
This result is also consistent with the model predic-
tion of Casey and Sposito Ž1992. for dissolution at
pH - pH pznpc , assuming that the pH pznpc of forsteritic
olivine equals 10.5 " 0.2 ŽWogelius and Walther,
1991. or 9.1 ŽSverjensky, 1994..
Our observation with respect to pH-dependence at
658C differs from that of Wogelius and Walther
Ž1992. who reported no significant change in the
pH-dependence from 258 to 658C based on a compar-
ison between the low temperature data of Wogelius
and Walther Ž1991. and the data they reported at
658C. Their 658C data are also plotted in Fig. 6
Žopen triangles.. There are three possible reasons for
the discrepancy between our results at the higher
temperature and those obtained by Wogelius and
Walther Ž1992.. First, all of our experiments were
conducted in solutions of HCl only, while in
Fig. 6. pH-dependence of olivine dissolution at 258C and 658C. Wogelius and Walther’s experiments, a buffer solu-
The line running through the 258C data represents a linear regres- tion was used for one of their experiments Žfor
sion of data by Blum and Lasaga Ž1988. and Wogelius and pH s 1.8.. Note, however, that they show evidence
Walther Ž1991., without the data at pH 5.7. The line running
that the buffer solution has a very small effect on
through the 658C data is the regression of the data from this work
Žwithout the data at pH 5. normalized to initial surface area Ždata dissolution at pH 2. Second, all of our experiments
points shown are also normalized to initial surface area.. except the one at pH 5 show stoichiometric release
of Mg, Si and Fe, while in Wogelius and Walther’s
experiment at pH 6, dissolution was presumed incon-
Žnormalized by initial surface area. is constrained gruent. In our experiments, the data at pH 5, which
between y9.92 and y10.0 molrcm2rs, while the show non-linear dissolution, were not used in deter-
log rate measured by Wogelius and Walther Ž1992. mining the pH-dependence. Because few data points
at pH 1.8 lies between y10.1 and y10.5 were used by Wogelius and Walther Ž1992. to deter-
molrcm2rs, assuming their error equalled the size mine the pH-dependence, an error in one data point
of their symbols ŽWogelius, pers. comm... If we would tremendously change the pH-dependence. Fi-
assume that error in measurement of surface area nally, it is possible that the rate at pH 6 used in the
between laboratories may be larger than 10%, then regression by Wogelius and Walther Ž1992. lies in
these rate measurements may be well within error. the pH-independent part of the dissolution rate–pH
Comparison of data normalized to initial surface curve. In support of this, we note that the rate we
area obtained at the two temperatures Žour data and measured at pH 5 is consistent with the rate reported
data reported by Blum and Lasaga Ž1988. and by Wogelius and Walther Ž1992. at pH 6 Žsuggesting
Wogelius and Walther Ž1991.. suggests the pH-de- that rates above pH 5 lie in the pH-independent part
pendence of the dissolution rate, n, increases with of the curve.. For these reasons, in fitting data at
temperature. At the 95% confidence level, n mea- both 25 and 658C, we did not use the rates measured
sured at 258C Ž0.49 " 0.095. differs from n mea- for pH ) 5.
sured at 658C Ž0.70 " 0.065.. This observation is New data by Rosso and Rimstidt Žin press. also
similar to that for pyroxene and amphibole dissolu- are consistent with a model wherein n is constant
tion for which we observed an increase in the pH-de- with temperature for forsteritic olivine dissolution
Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281 277

below 458C. We note that at a 99% confidence level To calculate the activation energy for the rate
for our data, we cannot reject the hypothesis that the constant ŽEq. Ž1.., we use a form of the Arrhenius
slope at 658C is equal to the slope at 258C, as equation:
predicted by these workers and Wogelius and Walther
2.303 R k1
Ž1992.. More measurements over a broad range of
temperature are needed to evaluate the temperature
dependence.
Ea s 1
T2 y 1
T1
ž ž //
log
k2
Ž 7.

We call this the pH-independent activation energy


Extrapolating the data of Rosso and Rimstidt Žin
because it is calculated for the rate constant and not
press. to 658C predicts rates slower than our rates by
the rate.
about a half order of magnitude ŽRimstidt, D., pers.
The activation energy calculated from the rate
comm... We infer that surface area measurements
constant of rate model Eq. Ž1. is 30 " 4 kcalrmol.
may be responsible for some of these discrepancies.
Because this value is equivalent to the value that
For example, Rosso and Rimstidt Žin press. used
would be calculated if an Arrhenius plot were com-
nitrogen for their BET measurements. In tests in our
pleted using the rate constant defined in Eq. Ž1., we
laboratory, using a Micromeritics ASAP 2010 Sur-
quote this value. However, this Ea , dependent upon
face Area Analyzer on a Micromeritics aluminosili-
n and the extrapolated value at pH s 0, incorporates
cate standard, surface area measured using N2 as the
extrapolation error. For this reason, we summarize
adsorbate averaged ; 35% higher than surface areas
ŽTable 4. the value of activation energy calculated at
measured using Kr. Such an effect is also well-
different pH values within the range of our experi-
established in the surface area measurement litera-
ments.
ture. Since we used Kr, our surface areas would have
In comparing the low temperature data to our
been systematically smaller than those reported by
data, we note that a variety of different solutions
Rosso and Rimstidt Ži.e., if we had used the same
were used for the low temperature experiments Že.g.,
grain size., and our rates would then have been
see Wogelius and Walther Ž1991. for solution chem-
systematically larger. In addition, the olivine used by
istry; Na perchlorate was used in the work of Blum
Rosso and Rimstidt Žin press. contained talc, and the
and Lasaga Ž1988. and Ž1988, A. Blum, pers. comm...
surface area had to be corrected for this impurity.
Differences in solution chemistry presumably con-
Therefore, the rate results from our group and their
tribute to differences in measured rates. For example,
predictions for 658C are probably equal within error.
Sanemasa et al. Ž1972. report that the dissolution of
Note that Kr BET was used by both Blum and
forsteritic olivine increases in solutions in the order
Lasaga Ž1988. and Ž1988, A. Blum, pers. comm..,
of HClO4 - HCl - H 2 SO4 . Thus, our activation en-
and Wogelius and Walther Ž1991; 1992..
ergy must be considered an estimate until more
An activation energy Ž Ea . can be estimated based
measurements are completed in constant solution
on the experimental data using the following equa-
chemistry.
tion:
The value of Ea calculated from the rate constant
ln Ž r 1rr 2 . s EarR Ž 1rT2 y 1rT1 . Ž 6. Ži.e., at pH s 0. is larger than the Ea estimated by
Wogelius and Walther Ž1992.. These latter workers
where r 1 and r 2 are the dissolution rates at pH of
observed that n did not vary with T, and therefore,
interest and at temperatures T1 and T2 , respectively.
their activation energy would be identical regardless
R is the gas constant. We have previously referred to
of whether it were calculated using Eq. Ž6. or Eq.
this as a pH-dependent activation energy ŽChen and
Brantley, 1998. because, when n changes as a func-
tion of temperature, Ea calculated in this way will
also change with pH. In addition, this activation
Table 4
energy is an apparent activation energy calculated Calculated activation energy for olivine dissolution
from two values of rates, rather than an activation pH 0 1 2 3 4
energy calculated using rate constants determined at Ea 30 27 25 22 20
two or more temperatures.
278 Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281

Ž7.. Note that this activation energy is not dependent Without these cations, activation of the mineral sur-
upon pH because n does not vary with temperature. face may be easier Žlower activation energy., even
Using Eq. Ž7. and calculating the Ea for our data though the unaltered mineral was one of a class of
between pH 2 and 4, the range Ž20–25 kcalrmol, more highly polymerized silicates. However, the ex-
see Table 4. compares well with the value reported tent of the data is limited for the full range of
by Wogelius and Walther Ž1992.: 19 kcal " 2.5 minerals, and mineral dissolution rates must be mea-
kcalrmol. Both of these values are larger, however, sured over wider ranges in temperature to test this
than the value reported for pH 2.5 by Westrich et al. trend.
Ž1993. for six compositions of orthosilicate Žs 12–16 The activation energy measured for olivine in this
kcalrmol., as measured for a temperature range work is higher than the Ea of diffusion Ž Ea ; 5
from 258 to 458C. Indeed, reported activation ener- kcalrmol., consistent with a surface-controlled
gies for olivine dissolution vary considerably. For model for dissolution. However, some of the activa-
example, Ea was reported to equal 14–16 kcalrmol tion energies previously cited in the literature for
at pH 1 Žin HClO4 and H 2 SO4 solutions respec- olivine are low enough that transport limitation may
tively, Sanemasa et al. Ž1972.., 9.1 kcalrmol at pH have occurred. The lack of dependence of dissolution
3.25 Žin HCl q KCl, Grandstaff, 1980, 1986., 6.0–7.2 rate on stirring rate ŽTable 3. in our experiments is,
kcalrmol at pH 1 Žin HCl and H 2 SO4 respectively, however, consistent with a surface-controlled mecha-
van Herk et al. Ž1989.., 16 kcalrmol Žin H 2 SO4 , nism for dissolution of olivine under these condi-
Jonckbloedt, 1997., and 10.2 kcalrmol by Rosso tions.
and Rimstidt Žin press.. Differences in treatment of
olivine powder, temperature range of measurements, 6.2. Secondary phase precipitation
experimental protocol, data analysis, and differences
in solution chemistry may all contribute to these
As noted earlier, dissolution of forsteritic olivine
inconsistencies.
under the experimental conditions of this study is
The value of Ea is consistent with a trend of
stoichiometric with respect to MgrSi. In contrast,
decreasing activation energy of dissolution in HCl–
for dissolution at ; pH 1, Seyama et al. Ž1996.
H 2 O with increasing connectedness of silicate tetra-
report that X-ray photoelectron spectroscopy of the
hedra Žnumber of bridging oxygens per tetrahedron.
leached surface of forsteritic olivine reveals selective
as measured in our laboratory: the Ea for forsteritic
leaching of Mg 2q Žand replacement by Hq. and an
olivine Ž30 " 4 kcalrmol, this study. ) diopside Ž23
unchanged silicate anion structure. In our experi-
" 2 kcalrmol. ) anthophyllite Ž19 " 1 kcalrmol,
ments, we observe no such nonstoichiometric leach-
Chen and Brantley, 1998.. In addition, as discussed
ing of Mg. However, the release rate of Fe becomes
in Brantley and Chen Ž1995., although there is no
nonstoichiometric at pH 5. It is possible that, be-
good measurement for the Ea of dissolution of en-
cause the reactor was open to the atmosphere, at
statite available, an estimate is ; 19 kcalrmol,
higher pH values iron hydroxides precipitated or
showing that for the strict Mg endmembers, the
ferric iron adsorbed to the olivine surface. Following
measured Ea values also roughly decrease with con-
Wogelius and Walther Ž1992., we calculate the rate
nectedness. Furthermore, the activation energy for
of Fe oxidation at atmospheric oxygen and pH 5
albite Ž15.6 " 0.8 kcalrmol, Chen and Brantley,
from the following equation ŽDavison and Seed,
1997. is also lower than some of the less ‘‘con-
1983; Millero et al., 1987; as cited by Wogelius and
nected’’ structures. This observation is surprising in
Walther, 1992.:
that one would have a priori predicted that higher
connectedness would correlate with higher Ea . The R ox Ž molrl s . s 3.33 = 10 11 Fe Ž II . PO 2Ž OHy .
2

trend may be related to the observation that olivine


dissolves stoichiometrically whereas the inosilicates where wFeŽII.x is the molarity of ferrous iron, PO 2 is
first preferentially lose cations from the M1, M2, the partial pressure of oxygen in atm, and ŽOHy. is
M3, or M4 sites. Similarly, the feldspars also lose the activity of hydroxide in solution. As discussed by
cations from the surface during acid dissolution. Wogelius and Walther, this equation was originally
Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281 279

proposed by Davison and Seed Ž1983., and then the cantly from the hydrated feldspar surface even when
ŽOHy. dependence was extended to pH 5 by Millero that feldspar surface has been leached of alkali,
et al. Ž1987.. Although the rate model is quoted for alkaline earth, and aluminum cations.
258C, similar to the conclusions of Wogelius and
Walther, we calculate that, during the 5 h duration of
our experiment, only about 0.1% of the ferric iron 7. Conclusions
released during dissolution would oxidize at pH 5.
Therefore, we conclude that, even with a temperature
Eight experiments of forsteritic olivine dissolution
enhancement of precipitation, after several hours,
using batch reactors were conducted at 658C between
rather than a ferric precipitate forming, the mineral
pH 2 and 5. A strong pH-dependence of dissolution
surface itself is probably oxidized, inhibiting the rate
rate is observed under the experimental conditions.
of dissolution.
Comparing the literature data on forsteritic olivine
The effect of Fe oxidation on olivine dissolution
dissolution at 258C, we observed an increase in
was also reported by Wogelius and Walther Ž1992..
pH-dependence at higher temperature, which is con-
In their experiments on fayalite dissolution open to
sistent with the prediction of a surface protonation
the atmosphere at 258C at 2 - pH - 4.5, Wogelius
model ŽBrady and Walther, 1992; Casey and Sposito,
and Walther Ž1992. observed a continuous decrease
1992..
in dissolution rate, which they attributed to oxidation
Based on literature data for forsteritic olivine
or precipitation of surface Fe.
dissolution at 258C and our results at 658C, we were
able to estimate the pH-independent activation en-
6.3. Effect of solution Al on dissolution rate of
ergy of forsteritic olivine dissolution Ž30 " 4
forsteritic oliÕine
kcalrmol.. Because pH-dependence Ž n. changes with
temperature, however, Ea at a given pH Žfor example
Previous experiments on albite dissolution show
the activation energy of weathering at a given pH.
that the rate of albite dissolution is dependent not
will decrease with increasing pH. Comparing this
only on temperature and solution pH, but also on the
data with our previous data for inosilicate dissolu-
Al concentration in the solution ŽChou and Wollast,
tion, values of the Ea calculated for rate Eq. Ž1.
1985; Oelkers et al., 1994; Chen and Brantley, 1997..
decreased from the Mg-endmember orthosilicate
Dissolution rates of forsteritic olivine at 658C and
Žforsteritic olivine, 30 " 4 kcalrmol., to the Ca–Mg
pH 2.95, measured with and without Al, are plotted
endmember pyroxene Ždiopside, 23 kcalrmol., to the
vs. Al concentration in Fig. 4. ŽFor the experiment
Mg-endmember amphibole Žanthophyllite, 19
with 10 ppm Ž0.37 mmolrl. of Al ŽRun OV-6 in
kcalrmol..
Table 2., the solution was supersaturated with re-
Results also show that existence of Al in solution
spect to boehmite Ž g-AlOŽOH.. at 658C based on
does not affect the rate of forsteritic olivine dissolu-
thermodynamic calculations using SOLMINEQ.. The
tion significantly under the experimental conditions.
result shows that over a wide range of Al concentra-
This result may indicate that Al does not form strong
tion, the dissolution rate of forsteritic olivine is not
crosslinks to the unpolymerized surface of forsteritic
significantly affected by Al. The observed Al-inde-
olivine and thus does not retard the dissolution of
pendent dissolution of forsteritic olivine may be
this phase under the experimental conditions.
related to the silicate structure. Adsorption of Al on
the surface of forsteritic olivine, if it occurs, might
not form stable crosslinks between the isolated silica
tetrahedra, in contrast to the effect of Al on the Acknowledgements
dissolution of feldspars ŽChou and Wollast, 1985;
Casey et al., 1988; Oelkers et al., 1994; Chen and This research was funded through grants from the
Brantley, 1997.. Indeed, Fig. 4 is consistent with the National Science Foundation ŽEAR 9305141. and
suggestion that the structure andror the surface from the David and Lucile Packard Foundation to
charge of the hydrated olivine surface differs signifi- S.L.B. We acknowledge help from D. Voigt, L.
280 Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281

Stillings, and S. Yau, and excellent reviews from P. ŽEds.., Rates of Chemical Weathering of Rocks and Minerals.
Maurice and R. Wogelius. [J.D.] Academic Press, Orlando, FL, pp. 41–57.
Hellmann, R., 1994. The albite–water system: Part I. The kinetics
of dissolution as a function of pH at 100, 200, and 3008C.
Geochim. Cosmochim. Acta 58, 595–611.
References Jonckbloedt, R.C.L., 1997. Olivine dissolution in sulphuric acid at
elevated temperatures: implications for the olivine process, an
Berube, Y.G., DeBruyn, P.L., 1968. Adsorption at the rutile-solu- alternative waste acid neutralizing process. In: Vriend, S.P.,
tion interface: I. Thermodynamic and experimental study. J. Aijlstra, H.J.P. ŽEds.., Geochemical Engineering: Curent ap-
Colloid Interface Sci. 27, 305–318. plications and future trends. J. of Geochemical Exploration 62:
Blum, A.E., Lasaga, A.C., 1988. Role of surface speciation in the 337–346.
low-temperature dissolution of minerals. Nature 331, 431–433. Kharaka, Y.K., Gunter, W.D., Aggarwal, P.K., Perkins, E.H.,
Brady, P.V., Walther, J.V., 1992. Surface chemistry and silicate DeBraal, J.D., 1988. SOLMIN88: A computer program for
dissolution at elevated temperatures. Am. J. Sci. 292, 639–658. geochemical modeling of water–rock interaction. USGS Wa-
Brantley, S.L., Chen, Y., 1995. Chemical weathering rates of ter-Resources Invest. Report 88-4227, 120 pp.
inosilicate minerals. In: White, A.F., Brantley, S.L. ŽEds.., Luce, R.W., Bartlett, R.W., Parks, G.A., 1972. Dissolution kinet-
Chemical Weathering Rates of Silicate Minerals. Mineralogi- ics of magnesium silicates. Geochim. Cosmochim. Acta 36,
cal Society of America Short Course, v. 31, Mineralogical 35–50.
Society of America, Washington, DC, 119–172. Machesky, M.L., 1989. Influence of temperature on ion adsorp-
Casey, W.H., Sposito, G., 1992. On the temperature dependence tion by hydrous metal oxides, In: Melchior, D.C., Bassett,
of mineral dissolution rates. Geochim. Cosmochim. Acta 56, R.L. ŽEds.., Chemical Modeling of Aqueous System II. Am.
3825–3830. Chem. Soc. Symposium Series 416: 282–292.
Casey, W.H., Westrich, H.R., Arnold, G.W., 1988. Surface chem- Millero, F.J., Sotolongo, S., Izaguirre, M., 1987. The oxidation
istry of labradorite feldspar reacted with aqueous solutions at kinetics of FeŽII. in seawater. Geochim. Cosmochim. Acta 51,
pH s 2, 3, and 12. Geochim. Cosmochim. Acta 52, 2795– 793–801.
2807. Murphy, W.M., 1985. Thermodynamic and kinetic constraints on
Carroll-Webb, S., Walther, J.V., 1988. A surface complex reac- reaction rates among minerals and aqueous solutions. Ph.D.
tion model for the pH-dependence of corundum and kaolinite Dissertation, University of California, Berkeley, CA Žunpub-
dissolution rates. Geochim. Cosmochim. Acta 52, 2609–2623. lished..
Chen, Y., 1997. Temperature and ph-Dependence of Silicate Oelkers, E.H., Schott, J., Devidal, J.L., 1994. The effect of
Dissolution Rate at Acid pH. PhD Dissertation, Pennsylvania aluminum, pH, and chemical affinity on the rates of alumi-
State University, University Park, PA, USA Žunpub... nosilicate dissolution reactions. Geochim. Cosmochim. Acta
Chen, Y., Brantley, S.L., 1997. Temperature- and pH-dependence 58, 2011–2024.
of albite dissolution rate at acid pH. Chemical Geology 135, Rosso, J.J., Rimstidt, J.D., in press, A high resolution study of
275–292. forsterite dissolution rates, Geochim. Cosmochim. Acta.
Chen, Y., Brantley, S.L., 1998. Diopside and anthophyllite disso- Sanemasa, I., Yoshida, M., Ozawa, T., 1972. The dissolution of
lution at 258 and 908C and acid pH. Chemical Geology, in olivine in aqueous solutions of inorganic acids. Bull. Chem.
press. Soc. Jpn. 45, 1741–1746.
Chou, L., Wollast, R., 1985. Steady-state kinetics and dissolution Seyama, H., Soma, M., Tanaka, A., 1996. Surface characterization
mechanisms of albite. Am. J. Sci. 285, 963–993. of acid-leached olivines by X-ray photoelectron spectroscopy.
Davison, W., Seed, G., 1983. The kinetics of oxidation of ferrous Chemical Geology 129, 209–216.
iron in synthetic and natural waters. Geochim. Cosmochim. Siegel, D.I., Pfannkuch, H.O., 1984. Silicate mineral dissolution at
Acta 47, 67–79. pH 4 and near standard temperature and pressure. Geochim.
Dove, P.M., Crerar, D.A., 1990. Kinetics of quartz dissolution in Cosmochim. Acta 48, 197–201.
electrolyte solutions using a hydrothermal mixed flow reactor. Siever, R., Woodford, N., 1979. Dissolution kinetics and the
Geochim. Cosmochim. Acta 54, 955–969. weathering of mafic minerals. Geochim. Cosmochim. Acta 43,
Furrer, G., Stumm, W., 1983. The role of surface coordination in 717–724.
the dissolution of g-Al 2 O 3 in dilute acids. Chimie 37, 338– Stillings, L.L., Brantley, S.L., 1995. Feldspar dissolution at 258C
341. and pH 3: reaction stoichiometry and the effect of cations.
Furrer, G., Stumm, W., 1986. The coordination chemistry of Geochim. Cosmochim. Acta 59, 1483–1496.
weathering: I. Dissolution kinetics of delta-Al 2 O 3 and BeO. Stumm, W., Furrer, G., 1987. The dissolution of oxides and
Geochim. Cosmochim. Acta 50, 1847–1860. aluminum silicates: Examples of surface-coordination-con-
Grandstaff, D.E., 1980. The dissolution of forsteritic olivine from trolled kinetics. In: Stumm, W. ŽEd.., Aquatic Surface Chem-
Hawaiian beach sand. Proc. 3rd Intl. Symp. Water–Rock istry. Wiley, New York, 197–220.
Interaction, 72–74. Sverdrup, H.U., 1990. The kinetics of base cation release due to
Grandstaff, D.E., 1986. The dissolution rate of forsteritic olivine chemical weathering. Lund Univ. Press, 246 pp.
from Hawaiian beach sand. In: Colman, S.M., Dethier, D.P. Sverjensky, D.A., 1994. Zero-point-of-charge prediction from
Y. Chen, S.L. Brantleyr Chemical Geology 165 (2000) 267–281 281

crystal and solvation theory. Geochim. Cosmochim. Acta 58, Westrich, H.R., Cygan, R.T., Casey, W.H., Zemitis, C., Arbold,
3123–3129. G.W., 1993. The dissolution kinetics of mixed-cation orthosili-
van Herk, J., Pietersen, H.S., Schuiling, R.D., 1989. Neutraliza- cate minerals. Am. J. Sci. 293, 869–893.
tion of industrial waste acids with olivine — The dissolution Wogelius, R.A., Walther, J.V., 1991. Olivine dissolution at 258C:
of forsteritic olivine at 40–708C. Chemical Geology 76, 341– Effects of pH, CO 2 , and organic acids. Geochim. Cosmochim.
352. Acta 55, 943–954.
Varadachran, C., Barman, A.K., Ghosh, K., 1994. Weathering of Wogelius, R.A., Walther, J.V., 1992. Olivine dissolution kinetics
silicate minerals by organic acids: II. Nature of residual prod- at near-surface conditions. Chemical Geology 97, 101–112.
ucts. Geoderma 61, 251–268.

You might also like