You are on page 1of 12

Pergamon

Geochimica et Cosmochimica Acta, Vol. 65, No. 7, pp. 1059 1070, 2001 Copyright 2001 Elsevier Science Ltd Printed in the USA. All rights reserved 0016-7037/01 $20.00 .00

PII S0016-7037(00)00570-6

Are quartz dissolution rates proportional to B.E.T. surface areas?


JEAN-MARIE GAUTIER,* ERIC H. OELKERS, and JACQUES SCHOTT
Ge ochimie: Transferts et Me canismes (UMR 5563 CNRS-Universite Paul Sabatier), 38 rue des Trente-six Ponts, 31400 Toulouse, France (Received February 11, 2000; accepted in revised form August 25, 2000)

AbstractA single quartz powder was dissolved at 200C and 250C under far from equilibrium conditions in atmosphere-equilibrated deionized water during a sequential series of experiments performed over one year in a titanium open system mixed ow reactor. Scanning electron microscope (SEM) photomicrographs show that morphological changes of this powder were dominated by grain edge rounding and etch pit formation. Etch pit walls rapidly evolve into unreactive negative crystal faceted forms; etch pit density and diameter are essentially constant during the experiments, indicating that dissolution predominantly deepened rather than widened etch pits. Measured B.E.T. surface areas increase linearly with mass of quartz dissolved to a value 5.6 times greater than that of the initial powder during the course of the experiments. Nevertheless, measured 200C far from equilibrium mass normalized dissolution rates remained constant within experimental uncertainty. It is concluded that the bulk of the observed increase in B.E.T. surface areas during dissolution consisted of essentially unreactive etch pit walls which contribute negligibly to mineral dissolution. As similar etch pit development is common in natural systems, geometric rather than B.E.T. surface areas may provide a more accurate parameter for estimating dissolution rates. Copyright 2001 Elsevier Science Ltd
1. INTRODUCTION

Accurate mineral surface areas are essential to assess weathering rates of natural systems (Helgeson et al., 1984; White, 1995; Jo nsson et al., 1995; Hodson et al., 1996, 1997; Brantley et al., 1999). Although most interpretations of mineral dissolution/precipitation kinetics assume a direct dependence of rates on B.E.T. surface areas (e.g., Rimstidt and Barnes, 1980; Schott et al., 1981; Dove and Crerar, 1990; Berger et al. 1994; Gautier et al., 1994; Oelkers et al., 1994; Oelkers and Schott, 1995, 1999; Devidal et al., 1997), several studies suggest that this may not be the case (Walter and Morse, 1984; Holdren and Speyer, 1985; Anbeek, 1992; Hochella and Baneld, 1995). The goal of this study is to assess the connection between B.E.T. surface areas and quartz dissolution rates through a series of dissolution experiments performed on a single quartz powder. The purpose of this manuscript is to describe the results of this experimental study and apply them towards the improved understanding of reactive surface area evolution during dissolution processes. 1.1. Theoretical background Dissolution reactions signicantly change the morphology of mineral surfaces. For example, natural dissolution provokes etch pit formation in such diverse environments as soils, diagenetically developed secondary porosity, and hydrothermally altered rocks (Krinsley and Doornkamp, 1973; Wilson, 1975; Grandstaff, 1978; Berner and Holdren, 1979; Berner and Schott, 1982; Velbel, 1984, 1986). Insight into these morphologic changes can be made using Periodic Bond Chain (PBC) theory developed by Hartman and Perdok (1952, 1955). This theory classies mineral faces according to the number of non-colinear uninterrupted strong bond chains (PBCs) parallel

*Author to whom correspondence should be addressed (gautier@lucid. ups-tlse.fr). 1059

to each face. The three types of faces are: (1) F (at) faces which contain two or more PBCs; (2) S (stepped) faces which contain one PBC, and (3) K (kinked) faces which do not contain any PBC. The relative reactivity of the various crystal faces can be deduced using the idealized crystal surface representation commonly known as the Kossel crystal (Kossel, 1927). The Kossel crystal is composed of cubic structural units; each cubic structural unit comprises the stoichiometric composition of the crystal. The coordination number of the cubic structural units is 6, one for each cube face. Cubic structural units on the Kossel crystal surface can have from 1 to 5 bonds with other units of the crystal. Examples of each of these types of units are illustrated in Figure 1. The relative ease to remove a structural unit from the crystal surface decreases with the number of bonds it has with other units in the crystal. Structural units having only one or two bonds with other structural units in the crystal are highly reactive and thus transitory. Crystal faces of the Kossel crystal are generally comprised of structural units having either 3, 4, or 5 bonds with other crystal units. Figure 2 shows the F, S, and K faces of a Kossel crystal. K-faces are comprised exclusively of structural units having 3 bonds with other crystal units; S-faces are comprised predominantly of structural units having four bonds with other crystal units, those having only 3 bonds with other crystal units are located at the end of incomplete rows; F-faces are comprised predominately of structural units having 5 bonds with other crustal units, those having only 3 bonds with other crystal units are located at the corners of incomplete faces. Because dissolution is most rapid for those structural units having the fewest bonds with other crystal units, it follows that the relative dissolution rates of the faces are: K S F. In accord with this framework, two major morphologic changes occur during dissolution: grain rounding and negative crystal formation. Grain rounding is a consequence of the fact that crystal edges and corners are composed of structural units that have only 4 and 3 bonds, respectively, with other crystal units, and are thus relatively easy to remove (Hirth and Pound,

1060

J-M. Gautier, E. H. Oelkers, and J. Schott

Fig. 1. Various cubic structural unit types on a Kossel crystal surface.

1957, Cheng and Coller, 1987). This process is illustrated in Figure 3. The origin of negative crystals are mineral holes like pores, etch pits, cracks, or uid inclusion voids. The growth of mineral holes is equivalent to classic crystal growth (Burton et al., 1951, Nielsen, 1983; Baronnet, 1984; Boistelle, 1985; Bennema et al., 1999) as shown in Figure 4, initially present K and S surfaces will disappear, eventually leaving only F-faces. The resulting faceted hole corresponds to a negative crystal (Boistelle, 1985; Gratz et al., 1990, 1991; Gratz and Bird, 1993). As it is faceted with F-faces, negative crystal surfaces dissolve at a much slower rate than holes faceted with K or S-faces. The structural unit for quartz is the silica tetrahedron, which has a coordination number of 4. Tetrahedra at the quartz surface are bonded to either 1, 2, or 3 other silica tetrahedra of the quartz crystal. Those having only one bond with other silica tetrahedra are highly reactive and transitory; quartz surfaces are dominated by tetrahedra having either 2 or 3 bonds with other silica tetrahedra. The major crystal faces, prism and rhombe-

hedral faces are comprised essentially of tetrahedra bonded with 3 other silica tetrahedra, and therefore constitute F-faces; other quartz faces, such as basal planes, contain large number of tetrahedra bonded with only two other silica tetrahedra, and therefore constitute K-faces (Gratz and Bird, 1993). It follows from the discussion presented above that quartz dissolution will be characterized by 1) the rounding of crystal edges, and 2) the formation of negative crystals having prism and/or rhombehedral faceted faces. The degree to which these two morphologic changes occur during quartz dissolution is assessed below.
2. EXPERIMENTAL METHODS Natural quartz grains having an average size of 3 mm were crushed with an agate mortar and pestle. X-ray diffraction analysis revealed the presence of only -quartz diffraction peaks. ICP-MS and ICP-AES chemical analysis showed this sample to contain 99.5% SiO2. The sample was sieved to obtain a 50 to 125 m size fraction. The resulting powder was ultrasonically cleaned and repeatedly rinsed in deionized water to remove ne particles. Photomicrographs of the mineral surfaces just after ultrasonic cleaning can be seen in Figure 5. No ne particles are evident on these mineral surfaces. The grains were further cleaned by dissolution at 250C in a mixed ow reactor for 5 d. Fine particles were observed leaving the system during the rst 48 h of dissolution; no nes were observed after this elapsed time. Following the release of ne particles, the output silica solution concentration was measured as a function of time. An initial concentration decrease was observed, followed by attainment of steady state. The initial fast Si release suggests the presence of a disturbed layer on the initial powder attributable to sample preparation (Rimstidt and Barnes, 1980). After this cleaning, the powder solubility was measured in a closed titanium reactor at 250C in atmosphere equilibrated deionized water (pH 5.7 at 25C). A value of 7.33 103 mol/kg was found, which is in good agreement with the -quartz solubility of 6.40 103 mol/kg reported by Walther and Helgeson (1977). These observations indicate that most ne particles and disturbed surfaces produced by sample preparation were removed by the cleaning process. The specic surface area of the quartz powder was determined by krypton adsorption via the B.E.T. method (Brunauer et al., 1938), and by geometric surface area calculation. The geometric specic surface area Ageo was calculated by assuming that particles are smooth spheres according to

Fig. 2. The three types of Kossel crystal surfaces: F (at), s, (stepped), and K (kinked) faces (Hartman and Perdok, 1955). The dashed cubes are bonded with only 3 other cubes of the Kossel crystal.

Are quartz dissolution rates proportional to B.E.T. surface areas?

1061

Fig. 3. Schematic illustration depicting the atomic-scale preferential dissolution of crystal corners.

Ageo

6 de

(1)

where de represents the effective spherical diameter and stands for quartz density (2.65 g/cm3). Assuming a homogeneous particle distribution leads to (Tester et al., 1994): de dmax dmin dmax ln dmin

(2)

where dmin and dmax refer to the minimum and maximum particle size, obtained from sieve aperture diameter. The computed values of de and Ageo for the ultrasonically cleaned powder are 82 m and 0.028 m2/g, respectively. Uncertainty associated with Ageo is difcult to assess owing to the unknown uncertainties in dmin and dmax. However, a reasonable value for these uncertainties is 10%, leading to an uncertainty of 10% for de and Ageo. The B.E.T specic surface area of the ultrasonically cleaned powder was ABET 0.061 m2/g, consistent with a roughness factor (Anbeek, 1992) of 2.2. Note that the Kr molecules adsorbed onto mineral surfaces to determine B.E.T. surface areas have an approximate diameter of 3.5. Therefore, un-like computed geometric surface areas, B.E.T. surface areas include contributions from small scale (1 nm) surface features. Dissolution experiments were performed at far from equilibrium conditions in the titanium mixed ow reactor system illustrated in Figure 6. Except the stirring system, this reactor is identical to that described by Berger et al. (1994) and Gautier et al. (1994). Inlet solutions consisted of doubly deionized atmosphere-equilibrated water having a pH of 5.7 at 25C. These inlet solutions were stored in a sealed low-density collapsible polyethylene (LDPE) container and en-

tered the reactor via a Gilson 307 HPLC pump tted with a 10WSC stainless steel head. The ow passed into a 215 mL titanium Parr pressurized reactor vessel kept at a constant temperature of 200 or 250 0.5C with a PID (Proportional Integral Derivative) controlled furnace. Outlet uid concentrations were measured regularly at 200C to obtain steady state dissolution rates. The temperature was increased to 250C during each experiment to accelerate the dissolution process, so that the effect of signicant dissolution could be observed. The powder in the reactor was stirred with a Parr titanium magnetic drive coupled to a stirring system designed to allow a good particle/ solution contact. To obtain a complete suspension of the 15 to 30 g of quartz powder used during the experiments, a newly designed large paddle was used to stir the solid and a bafe was attached to the reactor bottom (see Conti and Gianetto, 1986). The efciency of this system was validated by tests performed in a transparent reactor having dimensions similar to the 215 mL titanium reactor. Both the stirrer and the bafe were coated with several Teon layers to avoid powder abrasion. The uid left the reactor through a 2 m titanium lter and passes through a back pressure regulator. All the pH values reported in this study were determined using a XEROLYT electrode. The Si composition of efuent solutions was determined using the molybdate colorimetric method (Koroleff, 1976). 3. RESULTS AND DISCUSSION

3.1. Evolution of the quartz grain morphology during dissolution The evolution of grain surfaces was quantied using both B.E.T. surface area measurements and SEM observations of the

Fig. 4. Schematic illustration representing the disappearance of kinks during the negative crystal growth.

1062

J-M. Gautier, E. H. Oelkers, and J. Schott Table 1. Total grain mass variation ratios (total) and BET specic surface areas (ABET) determined in the present study. ABET (m2/g) 0.061 0.065 0.099 0.293 0.365 Experiment duration 200C (h) 0 139 240 492 0 1642 1686 2258 1720 Experiment duration 250C (h) 0 91 135 216 216 400 458 458 548

Powder Q0 Q1 Q2 Q3 Q4 Q5 Q6 Q7 Q8

total 0 0.0400 0.0644 0.0720 0.0722 0.196 0.211 0.224 0.247

Fig. 5. SEM photomicrograph of quartz grain surfaces after ultrasonic cleaning.

powders. Reaction progress during the sequential dissolution experiments can be quantied through the mass fraction of quartz dissolved. This mass fraction during any of the sequential experiments () was calculated using 60q Si t m (3)

where q represents the solution ow rate, [Si] refers to the silica concentration difference between the outlet and inlet at steady state, t corresponds to the experiment duration, m designates the mass of the powder before the experimental series, and 60 is the SiO2 molecular weight. The total mass fraction of quartz dissolved after any number of sequential dissolution experiments (total) can be computed from the sum of the values for all prior experiments. The values of total and the corresponding measured B.E.T. specic surface areas are listed in Table 1. The duration of each experiment is also given in this table. Q0 refers to the initial powder just after ultrasonic cleaning; Q1 to Q8 correspond to this powder after each sequential dissolution experiment. The

Fig. 6. Schematic illustration of the mixed ow reactor used in this study.

Are quartz dissolution rates proportional to B.E.T. surface areas?

1063

Fig. 7. Variation of quartz powder B.E.T. specic surface area as a function of total mass fraction of quartz dissolved.

variation of ABET as a function of total is depicted in Figure 7. The solid line represents a linear regression of the data obtained from powders Q1, Q4, Q6, and Q8. There is almost no ABET increase for total 4%, but ABET increases linearly at higher total to a value 5.6 times its initial value as total reaches 24.7%. The early anomalous behavior is likely due to linear arcuate etch pits germination (see below). Photomicrographs of the mineral surfaces at total 0.0400, 0.0720, and 0.224 are illustrated in Figure 8. Both the effects of grain rounding and etch pit formation are apparent in these photomicrographs. Grain rounding is evident by comparing Figures 5 and 8. Sharp edges and corners are observed on the starting grains; such features are rounded in all photomicrographs shown in Figure 8. Etch pit growth is apparent throughout the powder evolution. At the onset of dissolution (Fig. 8a) etch pits did not cover all grain surfaces and appeared as series of parallel linear arcs. Similar features were observed by Brantley et al. (1986) on quartz grains dissolved at 300C. These authors suggested that dislocations associated with surface scratches and micro-cracks may control these dissolution features. The most probable origin of such surface defects is sample preparation (Baronnet, 1988). As dissolution progressed, etch pits became rounded and were pervasive on all surfaces (Fig. 8b,c). A closer view of a typical rounded etch pit is shown on Figure 9. The polyhedral shape of this etch pit illustrates that it corresponds to a negative crystal. Etch pit density, estimated from SEM images, is estimated to be 108/ cm2. This value appears to be independent of total. Moreover, the average etch pit radius does not change during dissolution but remains close to 0.05 0.02 m. It follows that the B.E.T. surface area increase noted in Table 1 is likely due to etch pit deepening during dissolution. Similar surface area in-

creases have been noted for experimentally dissolved albite and diopside (Stillings and Brantley, 1995; Chen and Brantley, 1998). A simple model describing total surface area evolution during etch pit deepening can be developed by assuming (1) grains are spheres of constant radius, and (2) etch pits are cylinders of constant radius (i.e., etch pit walls do not contribute to quartz dissolution). Taking account of these assumptions leads to the relationship between the total mass fraction of quartz dissolved (total) and the specic surface area variation (A) given by (see appendix A) A 2 rb r total rn pit

(4)

where symbolizes the etch pit density, represents quartz density, rb stands for etch pit bottom surface area normalized dissolution rate, rn denotes the non-etch pit geometric surface area normalized dissolution rate, and rpit refers to the average etch pit radius. Equation 4 suggests that the total surface area increases linearly with the mass fraction of quartz dissolved. Such a linear relationship can be observed in Figure 7. Note that for cases where dissolution rates are time independent (e.g., steady state dissolution) Eqn. 4 implies that total surface area will increase linearly with time. The slope of the regression line in Figure 7 is 1.44 103 m2/kg; equating this slope rb with Eqn. 4 yields 12. This result indicates that etch pit rn bottoms dissolve at a surface area normalized rate that is 12 times faster than that of the non-etch pit surfaces. It should be noted, however, that because the total etch pit bottom surface area is two orders of magnitude smaller than that of the nonetch pit surface, it is the non-etch pit surface that dominates the

1064

J-M. Gautier, E. H. Oelkers, and J. Schott

Fig. 8. SEM photographs of typical quartz grains from (a) powder Q1 (total 0.0400); (b) powder Q3 (total 0.0720); and (c) powder Q7 (total 0.224).

overall dissolution process. This conclusion is consistent with the studies of mineral dissolution rates performed as a function of dislocation density by Casey et al. (1988) on rutile, Murphy (1989) on feldspar, and by Schott et al. (1989) on calcite. Each

of these studies observed only a minor variation in total surface area normalized dissolution rates despite the fact that dislocation densities, and presumably etch pit densities, varied by as much as ve orders of magnitude.

Are quartz dissolution rates proportional to B.E.T. surface areas?

1065

Table 2. Quartz dissolution rates measured at 200C in the present study. Mass normalized rate 107 (mol/kg/s) 1.59 1.88 2.28 2.02 Geometric surface area normalized rate 109 (mol/m2/s) 5.60 6.55 7.52 6.56 BET surface area normalized rate 109 (mol/m2/s) 2.45 1.90 0.778 0.553

Powder Q1 Q4 Q6 Q8

which decreased during the sequential dissolution experiments. Similarly, BET surface area normalized dissolution rates (rBET) at 200C were computed using rBET
Fig. 9. SEM photomicrograph showing a polyhedral etch pit on a quartz grain from powder Q2.

q[Si]s m ABET

(6)

where ABET refers to the BET specic surface area measured just before the experiment, and geometric surface area normalized dissolution rates (rgeo) at 200C were generated from rgeo q[Si]s m Ageo (7)

3.2. The variation of dissolution rates with B.E.T. surface area Steady state outlet Si concentrations were rapidly obtained during each sequential dissolution experiment. An example of a steady state outlet Si concentration is depicted in Figure 10. Mass normalized dissolution rates (rM) at 200C were computed from measured steady state outlet Si concentrations using rM q Si]s m (5)

where q represents the solution ow rate, [Si]s signies the output steady state silica concentration and m stands for the mass of powder present in the reactor during the experiment,

Fig. 10. Concentration of aqueous silica in the output solution of experiment conducted on powder Q5 as a function of elapsed time. The symbols in the Figure correspond to the measured silica concentration, and the solid line denotes the steady state concentration. The dashed vertical line represents the duration required to obtain 99% of steady state silica concentration (see Dove and Crearar, 1990).

where Ageo refers to the geometric specic surface area of the dissolving powder. Uncertainties in measured values of q is 1%. Uncertainty in [Si]s is on the order of 10%. This value corresponds to the reproducibility of output concentration measurements. Uncertainty in m, ABET and Ageo are on the order of 1%, 10% and 10%, respectively. Because total uncertainty in the computed rates equal the sum of the uncertainties in q, then the overall uncertainty in rM, rBET, and rgeo are 12%, 22%, and 22%, respectively. Measured mass, BET surface area, and geometric surface area normalized dissolution rates are given in Table 2. The corresponding experimental conditions are given in Table 3. Values of rM are illustrated in Figure 11 as a function of measured B.E.T. specic surface area. Mass normalized dissolution rates varied only within a factor of 1.27, despite the fact that the B.E.T. surface areas increased by a factor of 5.6 during the sequential dissolution experiments. Note the variations in measured mass normalized dissolution rates are approximately equal to the estimated uncertainties. It is concluded that B.E.T. surface area increases do not result in signicant mass normalized dissolution rate increases. A similar observation was reported by Grandstaff (1978), who reported constant mass normalized olivine dissolution rates during an experiment during which the measured BET surface area increases by a factor of 7. Because geometric surface areas do not include contributions from growing etch pits, they increase only slightly during the experiments. Consequently, geometric surface area normalized dissolution rates remain essentially constant during the experimental series. Conversely, BET normalized dissolution rates decrease by a factor of 4.5 during the sequential experiments. These observations are consistent with the etch pit development model presented above which assumed that etch pit walls are non-reactive during dissolution. A potential explanation as to why etch pit walls are nonreactive during dissolution is that these surfaces rapidly take

1066

J-M. Gautier, E. H. Oelkers, and J. Schott Table 3. Summary of experimental conditions of quartz dissolution at 200C in deionized water (present study).

Powder Q1 Q4 Q6 Q8
a b

total 0.0400 0.0722 0.211 0.247

Ageoa (m2/g) 0.028 0.029 0.030 0.031

ABET (m2/g) 0.065 0.99 0.293 0.365

Powder mass (g) 31.2 30.0 29.4 16.3

Flow rate (ml/min) 2.97 3.98 2.99 2.99

Output [Si] (mol/kg) 1.00 104 8.51 105 1.34 104 6.62 105

Saturation Indexb 0.027 0.023 0.36 0.018

pH-C 5.6020 5.3625 5.6217 5.5819

Values corrected for spherical grains size decrease according to Ageo 0.028 (1 total)1/3. Values computed with an equilibrium concentration of 3.74 103 mol/kg Walther and Helgeson, 1977).

the form of negative crystals which contain only relatively unreactive F-facets. Another potential explanation is that uid stagnation in small etch pits could result in a saturation effect, leading to smaller etch pit dissolution rates (Walter and Morse, 1984; Hochella and Baneld, 1995). However, calculations presented in Appendix B demonstrate that uid stagnation in quartz etch pits, which have radii of 5 108 m, do not result in signicant uid saturation. The bottom of etch pits, however, remain highly reactive thoughout dissolution. This may be because etch pits follow line defects present in the original quartz grains (see Nielsen and Foster, 1960; Spencer and Haruta, 1966; Lasaga and Blum, 1986; Brantley et al., 1986; Blum et al., 1990). Note some etch pits could be centered around point defects or microprecipitates (Baronnet, 1988; Blum et al., 1990) which will not signicantly deepen (Brantley et al., 1986). The density of active etch pits may therefore be less than the total observed. The discussion and observations presented above suggest that geometric rather B.E.T. surface areas may be more appro-

priate for describing mineral dissolution. This conclusion is consistent with experimental results reported by Tester et al. (1994). These authors prepared four powders from the same quartz sample, each having a distinct particle size. The far from equilibrium geometric surface area normalized dissolution rates of these powders are reproduced together with computed geometric surface areas in Table 4. It can be seen from this table that these rates varied randomly about a constant 5.7 1.5 108 mol/ m2/s, although the geometric surface area increased by a factor of 13. This observation indicates that the mass normalized dissolution rates are directly proportional to the geometric surface area. Similarly, recent work shows that geometric estimates of surface area can be used to accurately compute the overall dissolution rates of natural sandstones (Kieffer et al., 1999; Jove et al., 2001).
CONCLUSIONS

Steady state quartz dissolution rates were measured in atmosphere equilibrated deionized water at a temperature of 200C

Fig. 11. Measured mass normalized dissolution rate of quartz powders at 200C in deionized water as a function of B.E.T. specic surface area. The solid line is a linear regression of the experimental data.

Are quartz dissolution rates proportional to B.E.T. surface areas? Table 4. Quartz powder far from equilibrium dissolution rates determined at 200C in deionized water by Tester et al. Particle size (m) 74149 5374 149210 7071000 Geometric surface area (m2/g) 0.0211 0.0360 0.0127 0.00268 Geometric surface area normalized rate (mol/m2/s) 3.245.13 108 8.91 108 5.01 108 6.17 108

1067

and far from equilibrium conditions. The grains lost up to 24.7% of their initial mass during sequential dissolution experiments. SEM images reveal etch pit development on the grain surfaces as dissolution progresses, leading to signicant measured B.E.T. surface area increases. Nevertheless, little change was observed in measured mass normalized dissolution rates. The most probable explanation for this observation is that etch pit walls constitute the bulk of the surface are increase, but are relatively unreactive negative crystal facets. Consequently much of the B.E.T. measured surface area is unreactive. In contrast, the non-etch pit grain surfaces are rounded and may dissolve somewhat faster as dissolution progresses due to the creation of steps and kinks at surface edges and corners. Nevertheless as etch pit formation and growth are commonly a dominant mineral dissolution mechanism, geometric surface areas may be a better parameter for estimating mineral dissolution rates in both experimental and natural systems. This observation has major implications for estimating long term dissolution rates of natural soils and minerals. Although surface areas may increase dramatically during weathering (e.g., White, 1995), this additional surface area is likely unreactive and will contribute negligibly to the overall dissolution rate.
AcknowledgmentsWe are very grateful to Alain Baronnet and Olivier Grauby for many helpful discussions on crystal growth theory and practice. We would like to thank Gilles Berger, Olivier Jaoul, Bernard Moine and Jean-Louis Dandurand for useful discussions during the course of this study. We are grateful to the Mark Hodson and Johnathan Icenhower for their helpful reviews. We are indebted to Jean-Claude Harrichoury, Jocelyne Escalier, Pierre Brunet and Michel Gougeon for technical assistance. Supports from Centre National de la Recherche Scientique, Institut National des Sciences de la Terre, MESR fellowships and Universite Paul Sabatier are gratefully acknowledged. Associate editor: J. D. Rimstidt REFERENCES Anbeek C. (1992) Surface roughness of minerals and implications for dissolution studies. Geochim. Cosmochim. Acta 56, 14611469. Baronnet A. (1982) Ostwald ripening in solution. The case of calcite and mica. Estudios geol. 38, 185198. Baronnet A. (1984) Growth kinetics of the silicates. A review of basic concepts. Fortschritte der Mineralogie 62, 187232. Baronnet A. (1988) Mine ralogie. Coll Ge osciences (dunod), 184 p. Bennema P., Meekes H., and Van Enckevort W. J. P. (1999) Crystal Growth and Morphology: A Multi-Faceted Approach. In Growth, Dissolution and Pattern Formation in Geosytems. (eds. Bjorn Jamveit and Paul Meakin), Kluwer Academic Publishers, 21 64. Berger G., Cadore E., Schott J., and Dove P. M. (1994) Dissolution rate of quartz in lead and sodium electrolyte solutions between 25 and 300C. Effect of the nature of surface complexes and reaction afnity. Geochim. Cosmochim. Acta 58, 541551. Berner R. A. and Holdren G. R. (1979) Mechanism of feldspar weath-

ering. II: Observations of feldspars from soils. Geochim. Cosmochim. Acta 43, 11731186. Berner R. A. and Schott J. (1982) Mechanism of pyroxene and amphibole weathering II. Observations of soil grains. Am. J. Sci. 282, 1214 1231. Blum A. E., Yund R. A., and Lasaga A. C. (1990) The effect of dislocation density on the dissolution rate of quartz. Geochim. Cosmochim. Acta 54, 283297. Boistelle R. (1985) Concepts de cristallisation en solution. In Actualite s Ne phrologiques, Crosnier J. et al. (eds. Flammarion Me decine Sciences) 159 202. Brantley S. L., White A. F., and Hodson M. E. (1999) Surface area of primary silicate minerals. In Growth, Dissolution and Pattern Formation in Geosytems. (eds. Bjorn Jamveit and Paul Meakin), Kluwer Academic Publishers, 291326. Brantley S. L.,Crane S. R., Crerar D. A., Hellmann R., and Stallard R. (1986) Dissolution at dislocation etch pits in quartz. Geochim. Cosmochim. Acta 50, 2349 2361. Brunauer S., Emmett P. H., and Teller E. (1938) Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 60, 309 319. Burton W. K., Cabrera N., and Frank F. C. (1951) The growth of crystals and equilibrium structure of their surfaces. Phil. Trans. Roy. Soc. London 243, 299 358. Casey W. H., Carr M. J., and Graham R.A. (1988) Crystal defects and the dissoltuion kinetics of rutile. Geochim. Cosmochim. Acta 52, 15451556. Chen Y. and Brantley S. L. (1998) Diopside and anthophyllite dissolution at 25C and 90C and acid pH. Chem. Geol. 147, 233248. Cheng V. K. W. and Coller B. A. W. (1987) Monte Carlo simulation study on the dissolution of a train of innitely straight steps and of an innitely straight crystal edge. J. Crystal Growth. 84, 436 454. Conti R. and Gianetto A. (1986) Solid agitation and mixing in aqueous systems. In Encyclopedia of Fluid Mechanics, Vol. 2, Dynamics of Single-Fluid Flows and Mixing, 887900. Devidal J. L., Schott J., and Dandurand J. L. (1997) An experimental study of kaolinite dissolution and precipitation kinetics as a function of chemical afnity and solution composition at 150C, 40 bars and pH 2, 6.8 and 7.8. Geochim. Cosmochim. Acta 61, 51655186. Dove P. M. and Crerar D. A. (1990) Kinetics of quartz dissolution in electrolyte solutions using a hydrothermal mixed ow reactor. Geochim. Cosmochim. Acta 54, 955969. Gautier J. M., Oelkers E. H., and Schott J. (1994) Experimental study of K-feldspar dissolution rates as a function of chemical afnity at 150C and pH 9. Geochim. Cosmochim. Acta 58, 4549 4560. Grandstaff D. E. (1978) Changes in surface area and morphology and the mechanism of forsterite dissolution. Geochim. Cosmochim. Acta 42, 1899 1901. Gratz A. J. and Bird P. (1993) Quartz dissolution: Negative crystal experiments and a rate law. Geochim. Cosmochim. Acta 57, 965 976. Gratz A. J., Bird P., and Quiro G. B. (1990) Dissolution of quartz in aqueous basic solution, 106 236C: Surface kinetics of perfect crystallographic faces. Geochim. Cosmochim. Acta 54, 29112922. Gratz A. J., Manne S., and Hansma P. (1991) Atomic Force Microscopy of atomic-scale ledges and etch pits formed during dissolution of quartz. Science 251, 13431346. Hartman P. and Perdok W. G. (1952) A theory of crystal morphology. Kon. Nederl. Akad. Wet. Proc. B55, 134 136. Hartman P. and Perdok W. G. (1955) On the relations between crystal structure and morphology I, II, III. Acta Cryst. 8, 49 52.; 521524; 525529. Helgeson H. C., Murphy W. M., and Aagaard P. (1984) Thermodynamic and kinetic constraints on reaction rates among minerals and aqueous solutions. II. Rate constants, effective surface area, and the hydrolysis of feldspar. Geochim. Cosmochim. Acta 48, 24052432. Hirth J. P. and Pound G. M. (1957) Evaporation of metal crystals. J. Chem. Phys. 26, 1216 1224. Hochella M. F. and Baneld J. F. (1995) Chemical weathering of silicates in nature: A microscopic perspective with theoretical considerations. (eds. A. F. White and S. L. Brantley), Rev. Mineral. 31, Mineral Soc. Am., Washington, D.C, 353 406. Hodson M. E., Langan S. J., and Wilson M. J. (1996) A sensitivity

1068

J-M. Gautier, E. H. Oelkers, and J. Schott dissolution rates as a function of chemical afnity and solution composition. Geochim. Cosmochim. Acta 63, 785797. Oelkers E. H., Schott J., and Devidal J. L. (1994) The effect of aluminum, pH, and chemical afnity on the rates of aluminosilicate dissolution reactions. Geochim. Cosmochim. Acta 58, 20112024. Parks G. A. (1984) Surface and interfacial free energies of quartz. J. Geophys. Res. 89, 3997 4008. Rimstidt J. D. and Barnes H. L. (1980) The kinetics of silica-water reactions. Geochim. Cosmochim. Acta 44, 16831699. Robie R. A., Hemingway B. S., and Fisher J. R. (1978) Thermodynamic properties of minerals and related substances at 298.15 and 1 bar (105 Pascals) pressure and at higher temperatures. U.S. Geol. Survey Bull. 1452. Schott J., Berner R. A., and Sjo berg E. L. (1981) Mechanism of pyroxene and amphibole weathering. 1. Experimental studies of iron-free minerals. Geochim. Cosmochim. Acta 45, 21232135. Schott J., Brantley S.L., Crerar D., Guy C., Borcisk M., and Willaime C. (1989) Dissolution kinetics of strained calcite. Geochim. Cosmochim. Acta 52, 373382. Spencer W. J. and Haruta K. (1966) Defects in synthetic quartz. J. Appl. Phys. 37, 549 553. Stillings L. L. and Brantley S. L. (1995) Feldspar dissolution at 25C and pH 3: Reaction stoichiometry and the effect of cations. Geochim. Cosmochim. Acta 59, 14831496. Tester J. W., Worley W. G., Robinson B. A., Grigsby C. O., and Feerer J. L. (1994) Correlating dissolution kinetics in pure water from 25 to 625C. Geochim. Cosmochim. Acta 58, 24072420. Velbel M. A. (1984) Weathering processes of rock-forming minerals. Mineral. Assoc. of Canada Short Course Handbook. 10, 67111. Velbel M. A. (1986) Inuence of surface area, surface characteristics, and solution composition on feldspar weathering rates. In Geochemical Processes at Mineral Surfaces, Am. Chem. Soc. Symp. Series No. 323, 615 634. Walter L. M. and Morse J. W. (1984) Reactive surface area of skeletal carbonates during dissolution: Effect of grain size. J. Sed. Petrol. 54, 10811090. Walther J. V. and Helgeson H. C. (1977) Calculations of the thermodynamic properties of aqueous silica and the solubility of quartz an its polymorphs at high pressures and temperatures. Am. Jour. Sci. 277, 13151351. White A. F. (1995) Chemical weathering of silicates in nature: A microscopic perspective with theoretical considerations. (eds. A. F. White and S. L. Brantley), Rev. Mineral. 31, Mineral Soc. Am., Washington, D.C, 407 461. Wilson M. J. (1975) Chemical weathering of some primary rockforming minerals. Soil Sci. 119, 349 355.

analysis of the PROFILE model in relation to the calculation of soil weathering rates. Applied Geochemistry 11, 835 844. Hodson M. E., Lee M. R., and Parsons I. (1997) Origins of the surface roughness of fresh and unweathered alkali feldspar grains. Geochim. Cosmochim. Acta 61, 38853896. Holdren G.R. and Speyer P.M. (1995) pH dependent change in rates and stoichiometry of dissolution of an alkali feldspar at room temperature. Am. J. Sci. 285, 994 1026. Iler R. K. (1979) The Chemistry of Silica. Solubility, Polymerization, Collo d properties and Biochemistry. John Wiley and Sons Ed., New York, 866 p. Jo nsson C., Warfvinge P., and Sverdrup H. (1995) Uncertainty in predicting weathering rate and environmental stress factors with the PROFILE model. Water, air and soil pollution 81, 123. Jove C., Oelkers E. H., and Schott J. (in prep). An experimental investigation of long term variation of reactive surface area and permeability with dissolution of the Fontainebleau sandstone. Kieffer B., Jove C. F., Oelkers E. H., and Schott J. (1999) An experimental study of reactive surface area of the Fontainebleau sandstone as a function of porosity, permeability, and uid ow rate. Geochim. Cosmochim. Acta 63, 35253534. Koroleff F. (1976) Determination of silicon. In Methods of Seawater Analysis. (ed. K. Grasshoff), pp.149 158. Springer-Verlag. Kossel W. (1927) Zur Theorie des Kristallwachstums. Nachr. Ges. Wiss. Go ttingen. 135143. Krinsley D.H. and Doornkamp J. C. (1973) Atlas of Quartz Sand Surface Textures. Cambridge University Press. Lasaga A. C. and Blum A. E (1986) Surface chemistry, etch pits and mineral-water reactions. Geochim. Cosmochim. Acta 50, 23632379. Mizele J., Dandurand J. L., and Schott J. (1985) Determination of the surface energy of amorphous silica from solubility measurements in micropores. Surface Sci. 162, 830 837. Murphy, W. M. (1989) Dislocations and feldspar dissolution. Euro. J. Mineral. 1, 315326. Nielsen A. E. (1983) Precipitates: Formation, coprecipitation, and aging. In Treatise on Analytical Chemistry (eds. I. M. Kolthoff and P. J. Elving), pp. 249 347. Wiley. Nielsen J. W. and Foster F. G. (1960) Unusual etch pits in quartz crystals. Am. Mineral. 45, 299 310. Oelkers E. H. and Helgeson H. C. (1988) Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: Aqueous tracer diffusion coefcients of ions to 1000C and 5 kb. Geochim. Cosmochim. Acta 52, 63 85. Oelkers E. H. and Schott J. (1995) Experimental study of anorthite dissolution and the relative mechanism of feldpar hydrolisis. Geochim. Cosmochim. Acta 59, 5039 5053. Oelkers E. H. and Schott J. (1999) Experimental study of kyanite

Are quartz dissolution rates proportional to B.E.T. surface areas? APPENDIX A. MINERAL SURFACE AREA AS A FUNCTION OF GRAIN MASS VARIATION RATIO DURING ETCH PIT DEVELOPMENT R0 rb total 3rn

1069

(A11)

A model describing surface area evolution in response to etch pit development can be obtained by rst assuming grains are constant radius spheres, and etch pits are constant radius cylinders. The etch pit surface area increase (S1) resulting from a depth increase (h) is given by
S1 2 rpith (A1)

where rpit refers to the average etch pit radius. The total grain surface area increase (S) depends on etch pit surface area increase (S1) according to
S 4 R2 0 S1 (A2)

This last assumption (Sbrb Snrn) is equivalent to stating that the total dissolution stemming from etch pits is substantially less than that stemming from the bulk surface. This assumption is justied by the fact that the bulk surface is orders of magnitude larger than that of the etch pit bottoms. A relatively small enhancement of dissolution rates at etch pits is sufcient to create substantial etch pits with time. Hochella and Baneld (1995) suggested an enhancement of no more than 2 or 3 could create etch pits; the present study suggests this factor is about 12. Equation (A11) can be combined with Eqn. (A4) yielding
A 2 rb r total rn pit (A12)

where stands for the etch pit density, and R0 designates the grain radius. The grain specic surface area increase (A) as a function of S is given by
A S 4 R3 0 3 (A3)

This equation gives the increase in surface area as a function of total quantity of quartz dissolved.
APPENDIX B. EVALUATION OF SOLUTION SATURATION INDEX ADJACENT TO CYLINDRICAL ETCH PIT WALLS

where represents quartz density. Combining Eqn. (A3) with Eqns. (A2) and (A1) yields
A 6 rpith R0 (A4)

B.1. Calculation of the Si concentration prole within an etch pit cylinder As shown by Berger et al. (1994), overall quartz dissolution rates in pure deionized water obey a linear rate law; the overall quartz dissolution rate is the product of the far from equilibrium dissolution rate and the term (1 ), where designates the uid saturation index adjacent to the quartz surface. To assess if a saturation index increase could be responsible for the low etch pit walls dissolution rates observed in this study, aqueous silica concentration proles can be computed from the law of mass conservation. Aqueous silica can be added to solution by dissolution of either etch pit walls or bottoms. Assuming the uid is stagnant within etch pits, transport is diffusion limited. Mass conservation therefore requires
2 rpit D

Etch pit depth as a function of time can be estimated from


h M r rn t b (A5)

where M represents quartz molecular weight, rb stands for etch pit bottom surface area normalized dissolution rate, and rn denotes the non-etch pit geometric surface area normalized dissolution rate. The total number of moles of material dissolved from one grain is given by
n (Snrn Sbrb)t (A6)

where Sb and Sn represent respectively the sum of etch pit bottom surfaces and the total non-etch pits surface of one grain, which can be computed using

d2C C(x) 2 rpit rL 1 dx2 Cr

(B1)

2 Sb 4 R2 0 rpit

Sn 4 R2 0 Sb

(A7)

Combining Eqn. (A5) with Eqn. (A6) yields


h M rb rn n Snrn Sbrb (A8)

where rpit represents the etch pit radius, D denotes the aqueous silica diffusion coefcient, rL designates the etch pit wall surface area normalized far from equilibrium dissolution rate, C(x) symbolizes the silica concentration at the distance (x) from the etch pit bottom, and Cr represents the Si concentration in equilibrium with the etch pit wall. Taking account of GibbsThomson effects, Cr depends on the etch pit radius (rpit) according to (Iler, 1979; Baronnet, 1982; Mizele et al., 1985)
Cr Ceq exp

The total mass fraction of quartz dissolved is dened as


total Mn 4 R3 0 3 (A9)

2vm 1 R T rpit

(B2)

Combining Eqn. (A8) with Eqn. (A9) gives


h rb rn 4 R03 3 Snrn Sbrb

where Ceq symbolizes the Si concentration in equilibrium with an innite radius quartz particle, R represents the gas constant, T refers to the absolute temperature, designates the quartzwater interfacial tension, and vm denotes the quartz molecular weight. Eqn. (B1) can be further simplied to
d2C 2rL 2rL C(x) dx2 rpit DCr rpit D (B3)

total

(A10)

2 Assuming (1) Sb 4 R2 0, and thus Sn 4 R0, and (2) rb rn, but Sbrb Snrn, Eqn. (A10) reduces to

The boundary conditions for Eqn. (B3) are

1070

J-M. Gautier, E. H. Oelkers, and J. Schott

Fig. B1. Variation of solution saturation index adjacent to cylindrical etch pit walls as a function of distance from the etch pit top (see text).

C(x h) Ch D dC dx

(B4) (B5)

rb
x0

where h represents the etch pit depth, Ch refers to the Si concentration at the top of the etch pit, and rb stands for the etch pit bottom surface area normalized dissolution rate. Taking account of these boundary conditions, the general solution to Eqn. (B3) is
C(x) C1 exp(x) C2 exp(x) Cr (B6)

A value of Cr 2.18 mol/m3 is obtained from Eqn. (B2) using the parameters listed in Table B1. The values of (x), computed from Eqn. (B6), (B7) and the parameters in Table B1, are shown in Figure B1 as a function of distance from the etch pit top (i.e. the grain surface). The maximum value of the computed saturation index is 0.040. This degree of saturation will have negligible effects on overall etch pit wall dissolution rates.

where

2rL rpitDCr rb

ChCr , C1

rb

D , exp(h)exp(h)

exp(h)

Table B1. Parameters used to evaluate the degree of saturation adjacent to etch pit walls in this study. vma (cm3/mol) 22.688 rpit (m) 5 108
d

ChCr and C2

D . exp(h)exp(h)

exp(h)

b (J/m2) 0.35

Ceqc (mol/kg) 3.74 103

wd (kg/m3) 871

De (m2/s) 1.68 108 hi (m) 82 106

B.2. Numerical evaluation of solution saturation index adjacent to an etch pit wall The solution saturation index adjacent to etch pit walls ((x)) is given by
(x) C(x) Cr (B7)

Chf rLg rbh (mol/m2/s) (mol/kg) (mol/m2/s) 9.63 105 6.56 109 7.87 108

a Robie et al. (1978); b Parks (1984); c Walther and Helgeson (1977); Water density at 200C and 100 bars reported in NBS/NRC Steam Tables, Hemisphere Publishing Co., N.Y., 1984; e Diffusion coefcient for Cl in water at 200C reported by Oelkers and Helgeson (1988); f Outlet Si concentration average value determined from Table 2; g Estimated by averaging the geometric surface area normalized dissolution rates listed in Table 3; h rb 12 rL; i Average grain diameter.

You might also like