You are on page 1of 11

Pergamon

Geochimicaet CosmochimicaActa,Vol. 61, No. 7, pp. 1475-1485. 1997 Copyright ) 1997 ElsevierScienceLtd Printed in the USA. All rights reserved 0016-7037/97 $17.00 + .00 P I I S0016-7037(97) 00024-0

Inhibition of calcite crystal growth by Mg 2+ at 100C and 100 bars: Influence of growth regime
MARC DELEUZE and SUSAN L. BRANTLEY Department of Geosciences, Pennsylvania State University, University Park, Pennsylvania 16802, USA
(Received November 21, 1995; accepted in revisedfi)rm Januar), 3, 1997)

A b s t r a c t - - F o l l o w i n g Shiraki and Brantley (1995), who found different growth mechanisms for calcite crystal growth at 100C and 100 bars total pressure, we investigated the inhibition of calcite crystal growth by Mg 2+ under the same conditions of temperature, pressure, and composition. The results showed that the Mg 2+ inhibition depends on the growth regime. For growth in the exponential rate regime, the growth of calcite is almost totally inhibited, and aragonite grows instead. For growth in the linear rate regime, both aragonite and calcite precipitate. On the other hand, for experiments best described by a parabolic growth model, Mg inhibits the growth rate of calcite, but no aragonite precipitates. The study also showed that Mg inserts into the calcite structure but not into the aragonite structure to any significant extent during growth. The aragonite growth, which is controlled by primary nucleation, suggests that neither nucleation nor growth are inhibited by Mg. Discrepant observations concerning Mg inhibition documented in the literature may be explained by such differences in the growth mechanism. Copyright 1997 Elsevier Science Ltd 1. INTRODUCTION The precipitation and the dissolution of carbonate minerals often control porosity and permeability of common lithologies (e.g., Schmidt and McDonald, 1979a,b; Lundegard and Land, 1986; Taylor, 1989; Weedman et al., 1992). In the past few years, the dissolution and the precipitation of calcite (CaCO3) has been well investigated by many authors (e.g., Berner and Morse, 1974; Morse, 1978; Plummer et al., 1978; Sj6berg and Rickard, 1984; Busenberg and Plummer, 1986; Compton and Unwin, 1990; Gratz et al., 1993). These studies have shown the importance of pH and Pco2 in calcite dissolution and precipitation, mainly at temperatures of 60C. Although the knowledge of calcite precipitation and dissolution rates under subsurface conditions would help to understand many natural phenomena, the data for calcite waterreaction rates at high temperatures are not numerous (e.g., Hirano and Kikuta, 1985; Higuchi et al., 1988; Talman et al., 1990; Beck et al., 1992). For these reasons, Shiraki and Brantley (1995) initiated a study of calcite growth at 100C and 100 bars total pressure and documented different growth regimes for precipitation (interpreted as linear, parabolic, or exponential growth rate models). Pointing out that linear, parabolic, and exponential growth rate models can be derived for adsorption, screw dislocation, and surface nucleation growth mechanisms, Shiraki and Brantley (1995) suggested that calcite growth at 100C and 100 bars occurred according to different mechanisms. Growth rate of calcite in all the regimes occurred relatively quickly. Considering that natural reactions take place in impure media, we have investigated calcite growth inhibition for surface-controlled precipitation under those different growth mechanisms in the presence of dissolved Mg. Crystallization of carbonate minerals has been found to be inhibited by some solutes such as magnesium, phosphate, heavy metals, and organic products (e.g., Morse, 1983). We hypothesized that the presence of inhibitors might significantly slow the rate of calcite crystallization, 1475 even under diagenetic conditions. Due to the ubiquity of seawater as pore fluid and the high concentration of dissolved Mg in seawater, we chose to investigate the effect of Mg 2+ on calcite precipitation under the same conditions (temperature, pressure, composition) adopted by Shiraki and Brantley ( 1995 ). Over the last century, the kinetic inhibition of carbonatewater-reactions by Mg 2+ have been well studied at low temperature. The presence of dissolved Mg favors the precipitation of CaCO3 as aragonite instead of calcite from supersaturated seawater and other Mg-rich aqueous solutions (Leitmer, 1910, 1916; Lippman, 1960, 1973; Kitano, 1962; Simkiss, 1964). Other workers (Taft, 1967; Bischoff and Fyfe, 1968) showed that the recrystallisation of aragonite into calcite is inhibited even for very low ratios of Mg/Ca (ratio of dissolved Mg to dissolved Ca). The first detailed study on aragonite and calcite precipitation was made by Berner (1975) who found that only calcite crystallization is inhibited by dissolved Mg. However, Berner ( 1975 ) reported that at Mg 2+ levels less than about 5% of seawater, Mg 2 does not appreciably retard the seeded precipitation of calcite. Whereas for levels up to 5%, dissolved Mg in seawater severely retards the rate of seeded precipitation of calcite. Katz (1973) showed that calcite precipitation is not stopped even at Mg concentration levels approaching that of ocean water, concluding that the Mg/Ca ratio is sufficiently low in ocean water to not suppress calcite precipitation. More recent studies (Reddy and Wang, 1980; Mucci and Morse, 1983b) show that calcite growth rate decreases with increasing dissolved Mg concentration. These relatively recent studies did not mention any evidence of aragonite precipitation as suggested by the earlier experiments. Therefore, the effect of Mg 2+ on calcite crystallization is still controversial at low temperature. Differences in experimental conditions between these studies (solution chemistry, temperature, pressure, saturation state) may explain discrepancies: however, no systematic study has been completed.

1476 2. RATE LAWS

M. Deleuze and S. L. Brantley

Shiraki and Brantley (1995) summarized both physical models of surface-controlled crystallization as well as mechanistic models based upon elementary rate laws. As the aim of this work is to investigate the effect of Mg 2+ on the crystal growth rate of calcite when the mechanism is surfacecontrolled, and to compare the results with the Mg-free experiments, we summarize the precipitation rate laws based upon physical models of crystal growth used by Shiraki and Brantley (1995) to model the precipitation of calcite a t 100C and 100 bars total pressure (Table 1). These laws, derived for models of surface-controlled crystallization limited by adsorption, spiral growth at screw dislocations, or by two-dimensional surface nucleation on crystal surfaces have been summarized by Nielsen (1983) and Mullin (1993). In each of the three rate models summarized in Table 1, Rppt is the growth rate (tool cm -2 s-~), k is the rate constant (tool cm -2 s ~), A G is the driving force of crystallization, R is the gas constant, T is the absolute temperature, and f2 is the saturation state (exp(AG/RT). Although Shiraki and Brantley (1995) did not prove actual physical mechanisms (e.g., adsorption vs. screw dislocation growth), they did observe different regimes of Pco2 and A G where the rate laws in Table 1 applied. As indicated in Table 1, the three models predict a linear dependence (adsorption), a parabolic dependence (screw dislocation growth), and an exponential dependence on A G (surface nucleation). Numerous authors have used these rate models to describe the precipitation of various crystals: e.g., quartz (Rimstidt and Barnes, 1980), silver chloride (Davies and Jones, 1955 ), kaolinite, and gibbsite (Nagy et al., 1990, 1991 ; Nagy and Lasaga, 1992). Blum and Lasaga (1987) used Monte Carlo calculations to show that for the spiral growth-limited mechanism, n = 2 - 3 describes growth controlled by screw dislocations. Gratz et al. (1993) have cautioned, however, that many of the assumptions implied in analytical theories of crystal growth of calcite may be untrue, suggesting that simple attribution of mechanism based upon the form of the rate law may lead to error. Therefore, one of our goals in this investigation was to determine whether calcite growth in the three growth regimes (linear, parabolic, exponential) showed differences in the extent or effect of Mg 2 inhibition, as would be expected if the three regimes represented different mechanisms. For example, if Mg 2+ adsorption on the calcite surface inhibited nucleation, as suggested by Bischoff and Fyfe (1968), we would predict that the rate of calcite growth in the exponential growth regime would be decreased if such growth reflects surface nucleation. In contrast, if the parabolic growth regime corresponds to spiral growth at

Fig. 1. Scanning electron photomicrograph of calcite seed crystal.

screw dislocations, then no nucleation is required, and Mg 2+ adsorption might have only a limited effect. Finally, if Mg 2~ dehydration on the calcite surface is slower than Ca 2+ dehydration, then Mg might also inhibit calcite growth in the adsorption regime. 3. EXPERIMENTAL The following experiments reproduce exactly the conditions used by Shiraki and Brantley (1995).
3.1. Materials

Reagent grade CaCO3 (Fisher Scientific and Co.; Fig. 1) with a specific surface area of 0.175 _+ 0.011 m 2. g-~ (measured by 2 point Kr gas BET) was used as seed crystal for precipitation experiments. Note that some of the seed crystal exhibited rounded corners, presumably due to powder production processes. NaOH solutions for preparing as nutrient were first bubbled with CO2-Nz gas mixtures (see Table 2) for at least 24 h. The chemical compositions of solutions before the addition of Mg z were the same as those used by Shiraki and Brantley (1995) (Table 2). The saturation states of the solutions were adjusted so as to be undersaturated with respect to calcite at room temperature and supersaturated at high temperature. Solutions C2 and D, however, were slightly supersaturated with respect to calcite at room temperature. For these solutions, if the CaC12 and the MgCI2 were added before the bubbling, we observed after 24 h the precipitation of calcite in the nutrient solution reservoir. Therefore, CaC12 and MgCI2 were added just before the run to avoid any calcite precipitation in the nutrient solution reservoir. The concentration of Ca 2 and Mg 2 in the nutrient solutions were analyzed during each run and no change in composition was observed during the length of the experiments. Since the CaC12 and the MgCI2 were added individually for each experiment, small differences in Ca 2 and Mg 2+ concentrations are inevitable. Therefore, Table 2 lists the average concentrations for each solution. However, the differences are less than _+5%.
3.2. Runs

Table Mechanism

I : Growthmechanism and corresponding rate laws.


AG dependence Rate laws

Adsorption Spiral-growthat screw


dislocation

Li.... Parabolic

AG Rppt= k(exF(~)-1) AG n Rppt= k(exI(~)-l) ~ // RT)j Rppt =Aexn('\ k (AG

(1) (2) (3)

Surface-nucleation

Exponential

Crystal growth was performed in a continuously stirred tank reactor (300 mL volume, all wetted parts of Ti, Fig. 2). The solution was injected using a high pressure flow-monitoring pump, mixed completely throughout the vessel by a magnetically driven impeller, and allowed to flow out of the system through a back-pressure regulator. Effluent solutions were sampled periodically for Ca, Na, Mg, and pH analysis. The seed crystals were suspended in the vessel by stirring the solution with an impeller at 1200 rpm. Under these conditions, Shiraki and Brantley (1995) argued that calcite precipita-

Inhibition of calcite crystal growth by Mg 2+ Table 2 : Average compostion of nutrient solutions. Growth model Solutions pH Nat mM Ca2+ mM Mg"* mM CImM PCO2 TIC mM Alk mM Mg/Ca Parabolic B' 7.31 3.00 2.40 0.74 6.28 0.01 3.40 3.06
1/3

1477

B 7.31 3.13 2.05 0 4.10 0.01 3.43 3.09 0

B" 7.35 3.21

C2 7.12 6.45

Exponential C2' C2" 7.15 7.14 5.91 6.10 3.07 0.97 8.09 0.03 7.47 6.44
I/3

D 6.81 9.90

Linear D' 6.72 9.60 2.94 0.98 7.83 0.1 11.24 7.90
1/3

D" 6.74 9.27

2.05 1.38 6.86 0.01 3.40 3.09


2/3

3.10 0 6.20 0.03 7.37 6.30


0

3.11 1.90 10.02 0.03 7.29 6.27


2/3

3.01 0 6.02 0,1 12.92 9.59


0

2.85 1.79 9.28 0.I 11.71 8.30 2/3

tion was surface-controlled. Solutions were pumped through the reactor at room temperature for more than one hour before the heater was turned on. The run temperature was quickly obtained and a steady-state of Ca concentration reached. As calcite shows a retrograde solubility with temperature, the supersaturation at run temperature caused precipitation on the seed crystal. However, in some experiments (noted in results section), evidence of primary nucleation (Mullin, 1993), defined as spontaneous nucleation without seed, was identified because we saw sites of nucleation on the wetted parts of the reactor. Saturation state was varied by changing the amount of original seed crystal and by decreasing the flow rate during the experiment. The reactor was cleaned between each experiment by washing in dilute HCI. Figure 3 shows an example of the change in Ca concentration in the effluent solution as a function of time in the D ' series. As the solution was pumped through the reactor at room temperature for more than one hour before taking the first sample, the earliest measurements show the Ca concentration of the nutrient solution. After the reactor was heated, the steady state value of Ca concentration was quickly obtained. In this experiment, the fluid residence time in the reactor was approximately 0.9 h for a flow rate value of 9.5 10 5 L ' s t.
3.3. Solution Analysis

photometry). The pH of nutrient solutions was measured using a combination glass electrode before and during the run. The error ranges are _+2% for AAS and ICP and _+0.03 pH units for pH measurements. For nutrient solutions, the Total Inorganic Carbon (TIC) was calculated using the measured pH and Pco~ value of the bubbled gas. Alkalinity (Alk) was then calculated using the measured pH and the TIC value by (Morel, 1983) AIk - T I C / ( 1 0 Pu10~'-35 + 1) (1)

For a few runs, alkalinity was measured using a standard titration to pH - 4.5 in order to verify the calculation. The results obtained showed good agreement (_+3%) with the value calculated from the pH and the TIC. For effluent solutions, as the alkalinity can be calculated by AIk=[Na+]+2lCa2+]+2[Mg 2+] IC1 ] (2)

Na + concentrations in all solutions were measured by AAS (atomic absorption spectroscopy) whereas Ca and Mg concentrations were measured by ICP (Inductively Coupled Plasma spectro-

the alkalinity value for the effluent solution was calculated from measured Ca 2+ and Mg 2+ , and by assuming Na ' and CI concentration did not change during the run. The TIC was calculated by TIC~,i,~I-ACa-AMg, where ACa and A M g are the decrease of Ca 2+ and Mg 2- concentration, respectively, during the run. The pH value of effluent solutions was calculated from the TIC and the alkalinity, because the degassing observed in the sample tubing connected to the back pressure regulator did not allow use of measured pH values for the effluent solutions.
3.4. R u n Product Analysis

3.4.1. Magnesium analysis


In order to calculate the wt% Mg 2~ in calcite overgrowth, small amounts of run product were dissolved in a known volume of HNO~-

2.8 original seed: 98.1 mg 2.6 2.4 :~ 2.2 Flow rate 1:1.41E-4 '\ Flow rate 2:8.71E-5 \ I.s-1 I.s- 1

2
1.8

Fig. 2. Experimental apparatus, 1 : reaction vessel, 2: impellor and shaft, 3: magnetic stirrer, 4: solution inlet port, 5: filter (solution outlet), 6: control thermocouple, 7: pressure gauge, 8: mantle heater, 9: temperature controller, 10: temperature and stirring speed controller, 11: nutrient solution reservoir, 12:CO2-N2 gas mixture, 13: beaker with water, 14: high pressure flow-monitoring pump, 15:N2 gas (for total pressure control), 16: back pressure regulator, 17: cooling bath, 18: sample bottle.

1.6 1.4 O _ _ _
I J I

50
Time

100 (ran)

150

200

Fig. 3. Ca concentration in effluent solution as a function of time for solution D ' . Mass of original seed crystal and flow rates are indicated.

1478
Table 3 : Thermodynamicdata used. Reactions
H2C03 = H+ + H e 0 3 HCO 3 = H + + CO~CaHCO~ = Ca 2+ + H C O 3 C a C O3 = Ca 2+ + CO~-

M. Deleuze and S. L. Brantley of the wt% of each phase in our run products, within an acceptable accuracy (_+5%) over a range from 5 to 95% aragonite content. Since the amount of overgrowth material can be calculated from the decrease of Ca concentration and the duration of the experiment, the maximum theoretical wt% of aragonite can also be calculated assuming all the CaCO3 precipitation was aragonite. These results (listed in Table 4) can be compared to the measured wt% of aragonite in overgrowth material.
3.6. Saturation State

log K at 25C, 1 bar -6.35 -10,33 -1.ll


-3.22

log K at 100C, 100 bars -6.38 -10.|1 -1.00


-4.17

CaC03(s) =

Ca 2+

+ CO~-

-8.48 -13.99 -1.47

-9.17 -12.22 -1.91

1t20 = O H - + H +

co21g) + a 2 o

with

= a 2 c o "3

SOLMINEQ.88 (Kharaka et al., 1988) was used to calculate the speciation of solutions. The saturation index f~ was calculated from the product of the ion activities of calcium and carbonate ions (ac~ -~+, a c o ~ ) and the solubility product of calcite (K~p): ~), OCa2+ ac02 - -

H2CO 3 = C02(aq ) + H2CO3(aq )

Henry'slaw constant,KH

K~p

(3)

H20 solution and were analyzed by ICP. The amount of calcium carbonate precipitated in a given run was calculated from the decrease of Ca 2+ concentration and the duration of the experiment. Since the stirring speed was 1200 rpm for all experiments, we assumed an homogeneous distribution of initial seed and overgrowth material in the small amount of powder analyzed. Consequently, the measured values of Mg concentration were normalized and presented as the calculated wt% of Mg 2+ (Table 4) in overgrowth material. When the amount of precipitated material was sufficient, we checked the calculation by analyzing two samples from the same run. The % of Mg 2+ was always found to be the same within the error range.

3.5. X-Ray

Diffraction

Measurements

The run products were also analyzed by X-ray diffraction. In some experiments, both calcite and aragonite precipitation took place. In order to calculate the wt% of each phase, the areas under the most intense X-ray diffraction peaks of each product were measured. To normalize the results, we then determined XRD patterns for known quantities of mixed calcite and aragonite (ground together), measured under the same conditions. The empirical relation allowed calculation

However, Shiraki and Brantley ( 1995 ) noted that speciation calculations using the original SOLMINEQ.88 database showed that solutions in some runs were under-saturated with respect to calcite, although experimental observation indicated the solutions were at equilibrium. Talman et al. ([990) observed a similar inconsistency in their calcite dissolution experiments at 100, 150, and 210C, and they concluded that they needed to change the thermodynamic constants of the SOLMINEQ.88 database: they used log K, = - 1 for the association constant of CaHCO3+(aq) at 100C. This change in the database predicted f~ values close to unity for solutions which reached equilibrium. All other thermodynamic constants were assumed to be calculated correctly by SOLMINEQ.88. Table 3 lists the thermodynamic constants used for speciation of the carbonate system at 25C and 100C. Errors in saturation index were estimated in the same way as Shiraki and Brantley ( 1995 ), assuming the largest possible measurement errors. This approach yields a saturation state error of +12, 9, and 10% for B, C, and D series, respectively. 3.7. G r o w t h R a t e As the solution flows through the reactor at a constant rate, the precipitation rate can be determined using the decrease in Ca 2+

Table 4 : X-ray diffraction and ICP analysis results.

wt% Molality Solution aragonite ratio composition assuming Mg/Ca only aragonite
precipitation I

Decrease in

wt%
aragonite measured in

wt %
calcite measured in

calculated

wt% Mg in
precipitated material4

[Mg2 +] in effluent
solutions (%)

precipitate 2 7 6 30 22 0
0

precipitate 3 44 54 27 35 42
44

51 1/3 Linear
Model

1.6 1.2 1.3 1.4 1.3


1.4

3 3 3 3 5
5

sol D' 60 57

2/3 1/3
Parabolic

sol D" 57 42 sol B' 44 35 0 0 86 82 95 87 35 35 10 9 0 5 2 1.9 0.13 0.06 0.13 0,06 6.5 6.5 2 2 0 0

Model 2/3 sol B"

35 96 1/3
Exponential

sol C2' 91 94 sol C2" 92

Model 2/3 1 100 2 100 3 100 4 See

mass total precipitate/(mass precipitate + mass seed) mass aragonite precipitated/(mass precipitate + mass seed) mass calcite precipitated/(mass precipitate + mass seed) text

Inhibition of calcite crystal growth by Mg 2+ concentration between inlet and outlet. When a constant Ca 2+ concentration is reached in the effluent, material balance in terms of Ca yields
Rpp t =

1479

(C~.,~ - c~..o,)Q A,

(4)

where Cc,,m and Cc,,ou~are the nutrient and the effluent concentrations of Ca 2+, respectively, in moles L ~, Q is the flow rate in L s ~, Af is the final surface area in cm 2, and Rppt is the precipitation rate of the CaCO3 component of calcite in moles cm -2 s-~. In some experiments, incorporation of Mg 2+ into calcite overgrowth material was observed. Thus, to calculate the precipitation rate of all carbonate, Eqn. 4 should contain a term to account for the MgCO3 component. However, as we discuss later, this Mg 2+ uptake is very small (2%, Table 4) and thus, this term is very small. Consequently, Eqn. 4 expresses the rate of growth of the CaCO3 component in the precipitated carbonate, as well as the rate of growth of total carbonate within error. The final surface area is calculated using the relation AI = 0.034 mppt q- 0 . 1 7 9
mimsced

(5)

where mppt/mim~eedis the ratio of calcite mass precipitated during a run with respect to the mass of initial seed used. This relation was obtained by Shiraki and Brantley ( 1995 ) by measurement of surface area run products during growth of calcite under the same temperature and pressure conditions. 4. RESULTS Results of experiments are listed in Table 5. Table 4 shows a summary of the X-Ray diffraction measurements and the ICP analysis obtained for different growth regimes. Since Shiraki and Brantley (1995) showed that different growth models are applicable to different regimes of nutrient solution composition, we investigated the effect of Mg 2+ on the calcite crystal growth in each regime. As we discuss later, within each growth regime, the Mg 2+ effect is different. This effect is either manifested as a change in precipitation rate of calcite or as growth of aragonite instead of calcite. To compare the growth rate of Mg and Mg-free experiments, we plotted the precipitation rate of CaCO3 using the same models established by Shiraki and Brantley (1995). No attempt is made to identity the aragonite growth mechanism. 4.1. Linear G r o w t h E x p e r i m e n t s For calcite crystal growth in solution D, Shiraki and Brantley (1995) found that the data can be fit by the linear reaction rate equation
Rppt = 10 S,640.07(~'] __

on the wall of the reactor. The precipitated aragonite can be easily identified by SEM (Fig. 4a), because crystals are needlelike. In these experiments (Table 4), the Mg > analysis of run products contain between 1.2 and 1.6 wt% Mg 2" . This indicates a Mg uptake into the crystalline structure during growth. SEM photomicrographs of calcite also reveal microfractures (Fig. 4a and b). No fractures were observed in aragonite. Shiraki and Brantley ( 1995 ) did not observe such microfractures in calcite growth in their Mg-free experiments. Microfractures similar to those depicted in Fig. 4a,b have been previously documented by Berner ( 1975 ) on magnesian calcite precipited on calcite seed crystal. The ICP analysis of effluent solutions shows a maximum decrease of 3% in Mg > concentration between the inlet and outlet solutions in these experiments, which suggests Mg uptake into overgrowth material. Nevertheless, the ICP error ( _+2% ) does not allow precise evaluation of these results. In Fig. 5, we plot the growth rate of calcite against the saturation state (with respect to calcite) for solutions D, D ' , and D". The rate is decreased by the presence of Mg ( M g / Ca = 1/3 or M g / C a = 2/3). However, when the Mg > concentration increases from M g / C a = 1/3 to 2/3, the precipitation rate appears to increase slightly. This observation is probably not significant however. Because aragonite precipitates, the precipitation rate should actually be normalized by the surface area (Eqn. 4) of both calcite and aragonite rather than just calcite; therefore, an error occurs due to the great difference between calcite and aragonite surface area (0.175 m x g -t for calcite, Shiraki and Brantley (1995) and approximately 1.7 m 2 g ' for aragonite, Walter, pets. commun.). When the wt% of precipitated aragonite increases, more carbonate surface area is available for precipitation and the rate appears to be higher. In these runs. we did not measure the final surface area because the small quantity of overgrowth material (0.13 g maximum) did not allow precise measurement by B.E.T, and because some of this material was needed for ICP analysis. As shown in Fig. 5, the measured rate for both solutions D ' and D" are linear with respect to ~, as observed for similar solutions with M g / C a = 0 (Shiraki and Brantley, 1995). 4.2. Parabolic G r o w t h Experiments For solution B (without Mg2~), Shiraki and Brantley (1995) found that the growth mechanism is best described by a parabolic rate equation Rpp~ = 10 9.00+0.1~(~ l)l~=:l~ (7)

1 )1090.10

(6)

where Rppt is the precipitation rate of calcite (mol cm z s-~) and ~2 is the saturation state ( =exp ( A G / R T ) ) . Such a linear model can be derived for an adsorption growth model (Nielsen, 1983). The presence of Mg 2+ during experiments run with nutrient solution D leads to the growth of both calcite and aragonite (Table 4). Furthermore, the wt% of aragonite precipitated increases with increasing Mg concentration ( ~ 6 % for M g / C a = 1/3 and ~ 2 5 % for M g / C a = 2/3). The SEM observations (Fig. 4a) show that the aragonite can grow on calcite seed crystal, but that the major part of the precipitation takes place by primary nucleation (Mullin, 1993), defined as spontaneous nucleation without seed. The experimental observations also confirmed nucleation on a few sites

Such a parabolic growth model can be derived for crystallization rate-limited by growth at screw dislocations at low supersaturation (Burton et al., 1951 ). The results of our experiments using the same solutions but with dissolved Mg present are shown in Fig. 6. For M g / C a = 1/3, the growth rate of calcite is not affected by dissolved magnesium, whereas for M g / C a = 2/3, the precipitation rate shows a significant decrease. In these experiments, the X-ray diffraction measurements (Table 4) and SEM observations (Fig. 4e) show no aragonite. The ICP analysis of run products shows a slight difference between the two runs in terms of the amount of Mg 2+ incorporated during growth: 1.9 wt% Mg 2~ incorpo-

1480

M. Deleuze and S. L. Brantley


Table 5. Chemistry of effluent solutions and calculated rates.
Flow rate 1.41E-4 1.41E-4 1.41E-4 8.71E-5 8.71E-5 8.71E-5 4.60E-5 4.60E-5 1.41E-4 1.41E-4 1.12E-4 1.04E-4 8.33E-5 7.53E-5 1.41E-4 1.41E-4 8.71E-5 8.71E-5 4.60E-5 4,60E-5 1.42E-4 1.42E-4 9.86E-5 9.82E-5 1.41E-4 6,11E-5 5.39E-5 9.88E-5 6.14E-5 1.41E-4 1.40E-4 1.42E-4 1.12E-4 1.12E-4 9.03E-5 8.35E-5 8.29E-5 4.60E-5 1.41E-4 1.13E-4 1.41E-4 1.41E-4 1.12E-4 6.77E-5 1.12E-4 8.28E-5 8.36E-5 Ca mM 1.62 1.81 1.64 1.58 1.73 1.55 1.64 1.49 1.51 1.54 1.46 1.52 1.42 1.47 1.97 1.97 1.95 1.94 1.91 1.92 1,66 1,64 1.64 1.62 1.73 1.58 1.57 1.69 1.67 1.98 1.89 1.87 1.83 1.83 1.87 1.79 1.76 1.77 1.87 1.83 1.84 1.84 1.82 1.79 1.81 1.76 1.75 Mg mM 0.946 0.948 0.951 0.946 0.948 0.951 0.948 0.951 1,761 1.740 1.761 1.740 1.761 1.740 0.699 0.700 0.699 /).700 0.699 0.700 1.216 1.292 1.216 1.292 1.249 1.216 1.292 1.249 1.249 0.949 0.907 0.956 0.907 0.956 0.949 0.956 0.907 0.949 1.90 1.90 1.99 1.90 1.99 1.90 1.90 1.99 1.90 pH* 25C 6.44 6.42 6.43 6.42 6.39 6.40 6.37 6.39 6.44 6.43 6.43 6.42 6.42 6.41 6.78 6.77 6.76 6.74 6.73 6.72 6.85 6.83 6.83 6.81 6.83 6.78 6.78 6.79 6.78 6.64 6.63 6.62 6.61 6.60 6.59 6.58 6.58 6.55 6.58 6.57 6.56 6.56 6.55 6.55 6.54 6.52 6.52 pH* 100C 6.46 6.44 6.45 6.44 6.41 6.42 6.39 6.41 6.45 6.44 6.44 6.44 6.43 6.42 6.78 6.77 6.76 6.75 6.74 6.73 6,85 6,83 6.83 6.81 6.83 6.78 6.78 6.79 6.78 6.65 6.64 6.63 6.62 6.61 6.60 6.59 6.59 6.56 6.59 6.58 6.57 6.57 6.56 6.56 6.55 6.53 6.53 TIC* mM 10,27 10.01 9.99 10.23 9.98 9.90 9.89 9.84 10.68 10.37 1(I.63 10.34 10.59 10.29 2.96 2.90 2.94 2.87 2.91 2.85 3.42 3.39 3.39 3.37 3.(/3 3.33 3.32 2.98 2.96 6.20 6.41 6.36 6.35 6.32 6.09 6.28 6.28 5.99 5.97 5.93 6.21 5.91 6.20 5.90 5.87 6.14 5.82

Sol. D' D' D' D' D' D' D' D' D" D" D" D" D" D" B' B' B' B' B' B' B" B" B" B" B" B" B" B" B" C2' C2' C2' C2' C2' C2' C2' C2' C2' C2" C2" C2" C2" C2" C2" C2" C2" C2"

Alkalinity* mM
5.64 5.42 5.43 5.56 5.25 5.26 5.07 5.14 5.88 5.63 5.78 5.58 5.69 5.49 2.15 2.09 2.1 I 2.03 2.05 1.99 2.60 2.56 2.56 2.51 2.28 2.44 2.42 2.19 2.15 4.10 4.22 4.13 4.10 4.04 3.87 3.96 3.95 3.67 3.77 3.70 3.83 3.65 3.80 3.62 3.58 3.68 3.47

R m o l . c m -'-s ~ 6.00E-09 5.60E-09 5.30E-09 3,30E-09 3.24E-09 2.85E-09 1.55E-09 1.25E-09 6.67E-09 5.40E-09 4.50E-09 3.60E-09 3.10E-09 2.45E-09 3.80E-09 2.62E-09 2.22E-09 1.56E-09 1.11E 09 7.80E-10 2.60E-09 2.30E-09 1.80E-09 1.55E-09 1.30E-119 1.17E-09 9.90E- I 0 9.66E-10 6.11E-10 1.15E-9 1.31E-9 7.80E- 10 9.00E-10 4.5E-10 5.45E-10 2.3E- I 0 6.35E-10 2.45E-10 1.19E-9 7.05E- I 0 1.23E-9 I. 19E-9 7.49E-10 3.76E-10 6.85E-10 4.70E-10 4.49E-10

~LI 100C 0.50 0.52 0.41 0.36 0.31 0.21 0.16 0.12 0.38 0.31 0.30 0.28 0.21 0.19 0.59 0.52 0.48 0.37 0.34 0.28 0.82 0.69 0.69 0.59 0.60 (I.43 /).38 0.39 0.31 0,87 0.82 0.72 0.64 0.58 0.51 0.45 0.43 0.26 (/.53 0.44 /).43 0.40 0.38 0.33 0.29 0.25 0.17

* Calculated as described in text.

rated for Mg/Ca = 2/3, sol B" and 1.3 wt% tbr Mg/Ca = 1/3, sol B '. The ICP solution analysis also shows a small difference between the decrease of Mg 2+ concentration during the run: 5% for M g / C a = 1/3 and 6.5% for Mg/Ca = 2/3. Even if these differences are small with respect to measurement error, we argue that, in relation with the SEM observation, they are significant. Indeed, the run products show more microfractures when the Mg 2+ concentration is higher (Fig. 4 c - 4 e ) . This observation suggests that the Mg inserts into the calcite structure and produces growth defects. 4.3. Exponential G r o w t h E x p e r i m e n t s Shiraki and Brantley (1995) fit precipitation rate data for calcite growth in solution C2 for ~ - 1 > 0.6 to the following exponential model (Fig. 7):
Rppt = 10 7,280,49 e x p
2.36 -- 0.21

( AG/RT)

(8)

Such an exponential rate model can be derived for a surface-nucleation mechanism with mononuclear or polynuclear growth (Nielsen, 1983). In contrast, near equilibrium (i.e., $2 - 1 < 0.6), the precipitation is controlled by spiral growth as seen in solution B (Fig. 7). In attempts to measure calcite precipitation from solution C2 (Table 5) with dissolved Mg, we observed our results to be irreproducible. In crystallization experiments with solution C2 and M g / Ca = 1/3 or 2/3, X R D analysis of the run revealed high concentrations of precipitated aragonite. The X-Ray diffraction measurement shows that the wt% of aragonite is almost equal to the theoretical value calculated assuming that only aragonite crystallization occurred (Table 4). The fraction of aragonite present (based on X R D analysis) also indicates a difference between solutions C2' and C2": when the Mg 2+ concentration increases, the wt% of precipitated aragonite increases and approaches the theoretical value (Table 4). ICP analysis of run products show (Table 4) an average of 0.1% Mg 2+ in the precipitated product. We also observed

Inhibition of calcite crystal growth by Mg 2+ that the aragonite crystallization took place almost entirely on the reactor (walls and components, rather than on the seed) and that this effect increased with increasing Mg concentration. As seen before for solutions D ' and D", primary nucleation thus occurred during the experiment with C2' and C2", while for solution C2, surface nucleation controlled the growth. Because only aragonite crystallization took place, a plot of precipitation rate of calcite vs. saturation state of calcite is irrelevant. However, SEM observations reveal some information about the growth. Figure 4f shows calcite and aragonite crystals, and Fig. 4g shows growth of aragonite on calcite. In these photos, calcite crystals do not show any effect from the precipitation, in particular for the Mg/Ca = 2/3 experiments where we observed smooth surfaces without a lot of fractures (Fig. 4g, 4h). Also, in agreement with the low wt% of Mg observed in the run product (0.1%), we observed little fracturing in the calcite crystals, suggesting that almost no Mg 2+ is inserted in the calcite structure, but rather growth is almost totally stopped. Furthermore, there is little evidence that the Mg 2+ inserts into the aragonite structure. In support of this, the ICP solution analysis shows a decrease in Mg 2+ concentration of about 2% tbr Mg/Ca = 1/3 and no significant decrease for Mg/ Ca = 2/3. 5. DISCUSSION

1481

5.1. Influence of Mechanism


Summarizing our results, we observed: (1) When Mg 2+ is added to solutions where calcite grows by a linear model, both calcite and aragonite precipitate from the solution, and the wt% of aragonite in the precipitate increases with increasing Mg 2+ concentration. (2) When Mg 2+ is added to a solution in which calcite grows according to a parabolic model for Mg-free conditions, a sufficient amount of Mg 2+ (Mg/Ca = 2/3) reduces the precipitation rate of calcite and no aragonite precipitates. At lower concentrations of Mg 2+ (Mg/Ca = 1/3), the rate is not affected. (3) When Mg 2+ is added to a solution in which calcite grows according to an exponential model, the calcite growth is almost totally inhibited in favor of aragonite crystallization. Differences in the effect of Mg 2 on growth in each of the three regimes is consistent with different mechanisms of growth. These observations also yield implications with respect to the Mg 2~ inhibition mechanism, as discussed in the following. However, the arguments summarized below are qualitative in nature, reflecting the lack of firm observational evidence of mechanism and rate-limiting step.

Mucci and Morse, 1983) suggested that Mg :+ inhibits the rate of calcite nucleation. This model would suggest that growth rate-limited by surface nucleation should be strongly inhibited by the presence of Mg 2+ . If growth is rate-limited by surface nucleation for those experiments in the exponential growth regime (Nielsen, 1983), then our observation that calcite growth is almost totally inhibited in favor of aragonite growth in this regime is in agreement with these earlier authors. The process of inhibition is likely related to the smaller size, the higher charge density, and the resultant stronger hydration of Mg ~+ in comparison to Ca 2+. As the dehydration of reactants on the surface of a growing crystal is often a rate-controlling step in crystallization processes (Nancollas and Purdie, 1964), a hydrated Mg 2+ ion adsorbed on a calcite active growth site will remain for a relatively long time without dehydration and incorporation into the structure (in comparison with a Ca 2+ ion). This slow dehydration could prevent nucleation of calcite and lead to the crystallization of aragonite. A version of this model was first suggested by Bischoff (1968): as Mg 2+ inhibits calcite growth by blocking the nucleation, aragonite, which precipitates more rapidly, is kinetically stabilized. In other words, since calcite nucleation does not occur, the solution, which is still supersaturated with respect to calcium carbonate, cannot stay in a highly unstable state, and, therefore, aragonite precipitates. We also observe that for the experiments in the exponential growth regime, aragonite crystallized preferentially on the reactor walls, showing that, under these conditions, Mg 2+ inhibits neither nucleation nor growth of aragonite. The lack of an inhibitory effect by Mg 2+ on aragonite growth and nucleation can be explained by the observation that Mg 2+ is adsorbed at the surface of calcite to a much larger extent than at the surface of aragonite (Degroot and Duyvis, 1966), and also, because the Mg 2+ is not incorporated into the aragonite structure.

5.3. Linear Growth Model


For growth which is linear with respect to AG, growth may be rate-limited by adsorption on the surface structure and especially on the active site (Nancollas and Purdie, 1964). We observed that in calcite growth in the linear regime, Mg 2+ does not totally inhibit the calcite growth, but both calcite and aragonite crystallization takes place. This effect depends on the Mg 2+ concentration: the wt% of precipitated aragonite increases with increasing Mg 2+ concentration, suggesting that inhibition of calcite precipitation increases with Mg 2+ concentration. A hypothesis can be advanced to explain the observed results. As argued previously, the process of inhibition may reflect the stronger hydration of Mg 2+ in comparison to Ca 2+, since the dehydration of reactants on the surface of the growing crystal may be the rate-controlling step (Nancollas and Purdie, 1964). In this linear growth regime, if the process is controlled by adsorption, and if the dehydration of Mg 2+ is slower than that of Ca 2+, then the Mg 2+ adsorbed on the calcite structure may lead to inhibition of calcite growth, and, tbr the same reason as before, to aragonite crystallization. If inhibition depends on the density of cations adsorbed, then for low concentrations of Mg 2+ the probabil-

5.2. Exponential Growth Model


The orthorhombic structure of aragonite shows a strong preference for divalent cations the size of Ca 2+ or larger to insert into the structure (Speer, 1983). These structural arguments are in agreement with the observed lack of incorporation of Mg > in aragonite for solutions C2' and C2" (Table 4). The very low wt% of Mg 2+ (0.1%) in the run products tbr these experiments also confirms the (near-)absence of calcite growth, since Mg 2+ can be incorporated in calcite during growth. Previous workers ( Bischoff and Fyfe, 1968; Berner, 1975;

1482

M. Deleuze and S. L. Brantley

Fig. 4. SEM photomicrographs of run products. (a) Solution D' Mg/Ca = 1/3; (b) Solution D" Mg/Ca = 2/3; (c) Solution B' Mg/Ca = 1/3; (d) Solution B' Mg/Ca = 1/3; (e) Solution B" Mg/Ca - 2/3; (f) Solution C2' Mg/Ca = 1/3; (g) and (h) Solution C2" Mg/Ca - 2/3.

ity of occupying all the active sites is small, but this occupation will increase with increasing Mg z+ concentration. Furthermore, as calcite growth is not stopped by Mg > adsorption, the Mg 2+ can be incorporated into the calcite during precipitation. The insertion of Mg should produce some growth defects which could lead to the formation of microfractures during the growth. Finally, the effect of Mg 2+ is smaller in this growth regime than in the surface-nucleation growth regime because calcite nucleation still occurs (and is not rate-limiting), while growth of the calcite is inhibited.
5.4. Parabolic Growth Model

For precipitation in the regime of parabolic growth at low Mg concentrations (Mg/Ca = ~,(~),Mg 2+ is inserted into the calcite structure, but the precipitation rate is not affected. Thus, the data can be fit by the same law (Eqn. 7) used by Shiraki and Brantley ( 1995 ). This parabolic rate law is similar to the law attributed to growth at screw dislocations (Burton et al., 1951). For higher concentrations (Mg/Ca = 2/3), the precipitation rate shows a decrease, more Mg 2+ is incorporated into the precipitated calcite, and, the data are described by Rpp t = 1 0 - 9 5 1 + 1 4 ( ~ - 1 ) 1'33+015. For both

concentrations of Mg 2+, aragonite crystals do not grow from the solution. A possible explanation of these observations is found in the BCF theory for the spiral growth mechanism (Burton et al., 1951 ). A crystal possessing a screw dislocation contains a self-perpetuating step, and thus has no need for two-dimensional nucleation. For growth by such a mechanism, it is sometimes assumed that the surface diffusion of crystallizing species is the rate-controlling step (Nancollas and Purdie, 1964). On the other hand, Gratz et al. (1993) suggested that surface diffusion is not rate-limiting for step advance in growth of oxides and hydroxides in aqueous systems. Other rate-limiting steps might include volume diffusion in a boundary layer or a dehydration of adsorbed molecules at the surface. The dehydration of a Mg 2+ ion adsorbed on the surface, because of its high hydration energy, should be more difficult and slower than Ca 2+. For low concentrations of Mg the inhibitory effect may be expected to be negligible, and calcite crystallization should not be inhibited. However, at higher Mg 2+ concentrations, the probability of interaction between Mg 2+ and active sites is higher (even though diffusion on the surface is still slow) and the growth rate may be decreased. Thus, the inhibitory effect may be observed

Inhibition of calcite crystal growth by Mg-"

1483

Fig. 4. (Continued)

to be dependent upon Mg ~-+ concentration, as observed. Furthermore, as the crystal contains a self-perpetuating step which allows growth on a lot of active sites, and as in this case, the precipitation rate of calcite is low but not negligible (lower than the rate of the adsorption-controlled experiments), aragonite crystallization does not take place because there is no kinetic stabilization.

5.5. Implication of These Results


Although we have shown no direct evidence that the three growth regimes (linear, parabolic, exponential) correspond to the adsorption, screw dislocation, or surface nucleation mechanisms, the relative effects of Mg -~- inhibition observed are at least consistent with three mechanisms of growth. In

1.5 1 0 -9 Sol D Mg/Ca=0 Shirakl and Brantley ( Sol D' Mg/Ca=l/3 J ~ ~ -

4.0 1 0 - 1 A 3.0 1 0 "10 Sol B Mg/Ca=0 ~ Shiraki and Brantley (1995) Sol B' Mg/C8=1/3

N _~ o 5.o 10.10~ ~;~ ~ ~ ~ I

,w.

o E
n-

'E o 201 _~

0-1o

1.01

0-1o 0.8

0.1

0.2

0.3 fl-1

0.4

0.5

0.6

0.2

0.4 ~-1

0.6

Fig. 5. Calcite crystal growth rate plotted vs. ~-1 for solution D; D' and D" (linear growth). Saturation state (f~) calculated with respect to calcite using data of SOLMINEQ.88. Solution compositions are indicated.

Fig. 6. Calcite crystal growth rate plotted vs. ~-1 tor solution B, B ', and B" (parabolic growth). Saturation state (~) calculated with respect to calcite using data of SOLMINEQ.88. Solution compositions are indicated.

1484
2 . 0 1 0 "9
[2 0 Sol B Mg/Ca=O Shirakl and Brantley (1995)

M. Deleuze and S. L. Brantley

1.5 1 0 -9

Sol C2 Mg/Ca=O Shiraki and Brantley (1995)

E u ~ 1 . 0 1 0 -9 o E ~ 5 . 0 1 0 "1

/
0.8

/
1 1.2

0.2

0.4

0.6 fl-1

Fig. 7. Calcite crystal growth rate plotted vs. ft-1 for solution B and C2. Saturation state (f~) calculated with respect to calcite using data of SOLM1NEQ.88. Solution compositions are indicated.

the exponential growth regime, if calcite growth occurs by a surface nucleation mechanism, the observation that Mg 2+ almost totally inhibits calcite precipitation is consistent with a model whereby adsorption of Mg 2+ inhibits 2D-nucleation. Indeed, adsorption of Mg 2+ in solutions of low Mg/Ca has no effect on calcite growth where nucleation is assumed unnecessary: in the regime of parabolic (screw dislocation) growth. However, at higher values of Mg/Ca, even growth in this regime is inhibited, although the rate law of growth is still consistent with spiral growth at screw dislocations. Previous studies (see Introduction) based on natural media conditions have shown several discrepancies. Although our conditions (100C and 100 bars total pressure, dilute solutions) are very different from seawater under ambient conditions, our results may help to explain such discrepancies. If Mg 2+ inhibition depends on growth mechanism, discrepancies may be due to different growth regimes resulting in either an inhibition (partial or total) of calcite growth in favor of aragonite (Leitmer, 1910, 1916; Lippman, 1960, 1973; Kitano, 1962; Simkiss, 1964; Taft, 1967; Bischoff and Fyfe, 1968) or an inhibition of calcite precipitation without aragonite formation (Katz, 1973; Beruer, 1975; Reddy and Wang, 1980; Mucci and Morse, 1983b). For example, Berner (1975) reported spontaneous aragonite precipitation at the end of long runs for experiments investigating crystallization of magnesian calcite. The results of Berner (1975) might suggest a parabolic growth regime as he observed a low inhibition effect for low Mg 2+ concentration (5% of seawater) but a strong inhibition effect from solutions with Mg 2+ concentration equivalent to seawater. Nevertheless, the presence of aragonite crystallization only at the end of long runs cannot be explained. Thus, several discrepancies involving the effect of Mg on calcite growth and nucleation, over a range of temperature and pressure, remain to be investigated. 6. CONCLUSION The study of inhibition of calcite crystal growth by Mg 2. at 100C and 100 bars total pressure for surface-controlled experiments shows a strong influence of growth regime. The presence of low quantities of Mg (Mg/Ca = 1/3, 2/3) leads to Mg 2+ concentration-dependent phenomena: ( l ) Total

calcite growth inhibition in favor of aragonite crystallization in the exponential growth regime; (2) Partial calcite growth inhibition in favor of aragonite crystallization in the linear growth regime; (3) A decrease of the calcite growth rate without aragonite crystallization in the parabolic growth regime for Mg/Ca = 2/3; (4) No effect on calcite growth rate and no aragonite crystallization in the parabolic growth regime for Mg/Ca -- 1/3. Furthermore, these experiments also show that Mg > is incorporated in the calcite structure but not in the aragonite structure during nucleation and growth. The differences are consistent with different growth mechanisms operating within each regime. Since the strongest inhibition effect is observed for the growth regime where nucleation is hypothesized to control the rate of crystallization, Mg 2+ may inhibit calcite nucleation. This inhibition may be related to the high hydration energy of Mg 2+ and to the calcite structure which allows incorporation of Mg 2+ . On the other hand, neither nucleation nor growth of aragonite are affected by the presence of Mg 2+ ions because aragonite preferentially incorporates cations of ionic radii equal to or greater than that of Ca 2+, excluding small ions such as Mg 2+. These observations suggest that the presence of dissolved Mg in pore solutions under diagenetic conditions may significantly affect carbonate precipitation reactions in the subsurface. Depending upon the solution composition, calcite precipitation may be unaffected, may be slowed, or may be replaced by aragonite precipitation.
Acknowledgments--The authors would like to thank the DGA/

DRET for supporting M. Deleuze at Penn State, D. Voigt for help and discussion, H. Gong for the ICP measurements, C. Perry and M. Angelone for the SEM work, and R. Shiraki. Acknowledgment is made to the donor of the Petroleum Research Foundation,administered by the American Chemical Society, for partial support of this research.
Editorial handling." J. Tossell
REFERENCES

Beck J. W., Berndt M. E., and SeyfriedW. E., Jr. (1992) Application of isotopic doping techniques to evaluation of reaction kinetics and fluid mineral distribution coefficients:An experimental study of calcite at elevated temperatures and pressures. Chem. Geol. 97, 125-144. Bemer R. A. and Morse J. W. (1974) Dissolutionkinetics of calcium carbonate in sea water IV. Theory of calcite dissolution. Amer. J. Sci. 274, 1173-1186. Berner R. A. (1975) The role of magnesium in the crystal growth of calcite and araginote from seawater. Geochim. Cosmochim. Acta 39, 489-504. Bischoff J. L. and Fyfe W. S. (1968) The aragonite-calcitetransformation. Amer. J. Sci. 266, 65-79. Blum A. E. and Lasaga A. C. (1987) Monte Carlo simulations of surface reaction rate laws. In Aquatic SurJ2tce Chemist O, (ed. W. Stumm), pp. 255-292. Wiley. Burton W. K., Cabrera N., and Frank F. C. ( 1951 ) The growth of crystals and the equilibrium structureof their surfaces. Phil. Trans. Roy. Soc. London 243, 299-358. Busenberg E. and Plummer L. N. (1986) A comparative study of the dissolution and crystal growth kinetics of calcite and aragonite. U.S. Geol. Surv. Bull. 1578, 139-168. Compton R. G. and Unwin P. R. (1990) The dissolution of calcite in aqueous solution at pH < 4; Kinetics and mechanism. Phil. Trans. Roy. Soc. London A330, 1-45. Davies C.W. and Jones A.L. (1955) The precipitation of silver

inhibition of calcite crystal growth by Mg 2+ chloride from aqueous solutions. Part 2-Kinetics of growth of seed crystals. Trans. Faraday Soc. 51, 812-817. Degroot K. and Duyvis E. M. (1966) Crystal form of precipitated calcium carbonate as influenced by adsorbed magnesium ions. Nature 212, 183-184. Gratz A. J., Hillner P. E., and Hansma P. K. (1993) Step dynamics and spiral growth on calcite. Geochim. Cosmochim. Acta 57, 491 495. Higuchi M., Takeuchi A., and Kodaira K. (1988) Hydrothermal growth of calcite single crystals from H20-CO2-CaCO3 system. J. Crvst. Growth 92, 341-343. Hirano S. and Kikuta K. (1985) Hydrothermal crystal growth of calcite in NaCI and KC1 solutions. Chem. Lett., 1613-1616. Katz A. (1973) The interaction of Mg with calcite during crystal growth at 25-90C and one atmosphere. Geochim. Cosmochim. Acta 37, 1563- 1586. Kharaka Y. K., Gunter W. D., Aggarwal P. K., Perkins E. H., and De Braal J.D. (1988) SOLMINEQ.88: A Computer Program Code for Geochemical Modeling of Water-Rock Interactions. USGS Water Res. Investigation Report. 88, 4227. Kitano Y. (1962) The behavior of various inorganic ions in the separation of calcium carbonate from a bicarbonate solution. Bull. Chem. Soc. Japan 35, 1973-1980. Leitmer H. (1910) Zur Kenntnis der Carbonate: I. Die Dimorphie des Kohtensauren Kalkes. Neues. Jahrb. Mineral. I, 49-74. Leitmer H. (1916) Zur Kenntnis der Carbonate: II. Neues Jahrb. Mineral. Beilageband 40, 655-700. Lippmann F. (1960) Versuche zur Au f klarung der Bildungsbedingungen von Calcit und Aragonit. Fortschr. Mineral. 38, 156-161. Lippmann F. ~1973) Sedimentary Carbonate Minerals. SpringerVerlag. Lundegard P. D. and Land L. S. (1986) Carbon dioxide and organic acids: Their role in porosity enhancement and cementation, Paleogene ot: the Texas Gulf Coast. In Roles of Organic Matter in Sediment Diagenesis (ed. D.L. Gautier); Soc. Econ. Paleontol. Minelvl. Spec. Publ. 38, 127 146. Morel F.M.M. (1983) Principles of Aquatic Chemisto,. Wiley. Morse J. W. (1978) Dissolution kinetics on calcium carbonate in sea water. VI. The near equilibrium dissolution kinetics of calcium carbonate-rich deep sea sediments. Amer. J. Sci. 278, 344-353. Morse J. W. (1983) The kinetics of calcium carbonate dissolution and precipitation. Rev. Mineral. MSA Short Course 11, 227-264. Mucci A. and Morse J. W. (1983) The incorporation of Mg 2+ and Sr 2+ into calcite overgrowths: Influences of growth rate and solution composition. Geochim. Cosmochim. Acta 47, 217-233. Mullin J. W. ( 1993 ) Co'stallization. 3rd ed. Butterworth-Heinemann Ltd. Nagy K. L. and Lasaga A. C. (1992) Dissolution and precipitation kinetics of gibbsite at 80C and pH - 3: The dependence on solution saturation state. Geochim. Cosmochim. Acta 56, 30933111. Nagy K.L., Steefel C.I.. Blum A.E., and Lasaga A.C. (1990) Dissolution and precipitation kinetics of kaolinite: Initial results

1485

at 80C with application to porosity evolution in a sandstone. Amer. Assoc. Petrol. Geol. Mere. 49, 85-101. Nagy K. L., Blum A. E.~ and Lasaga A. C. ( 1991 ) Dissolution and precipitation kinetics of kaolinite at 80C and pH = 3: The dependence on solution saturation state. Amer. J. Sci. 291,649-686. Nancollas G. H. and Purdie N. (1964) The kinetics of crystal growth. Quart. Rev. Chem. Soc. London. 18, I. Nielsen A. E. (1983) Precipitates: formation, coprecipitation, and aging. In Treatise on Analytical Chemistry ( ed. 1. M. Kolthoff and P. J. Elving), pp. 269-347. Wiley. Plummer L. N., WigleT T. M. L., and Parkhurst D. L. (1978) The kinetics of calcite dissolution in CO2-water systems at 5 to 60C and 0.0 to 1.0 atm CO2. Amer. J. Sci. 278, 179-216. Reddy M. M. and Wang K. K. (1980) Crystallization of calcium carbonate in the presence of metal ions. 1. Inhibition by Mg at pH 8.8 and 25C. J. Crystal Growth 50, 470-480. Rimstidt J. D. and Barnes H. L. (1980) The kinetics of silica-water reactions. Geochim. Cosmochim. Acta 44, 1683-1699. Schmidt V. and McDonald D.A. (1979a) The role of secondary porosity in the course of sandstone diagenesis. Soc. Econ. Paleontol. Mineral. Spec. Publ. 26, 175 207. Schmidt V. and McDonald D. A. (1979b) Texture and recognition of secondary porosity in sandstones. Soc. Econ. Paleontol. Minerol. Spec. Publ. 26, 209-225. Shiraki R. and Brantley S. L. (1995) Kinetics of near-equilibrium calcite precipitation at 100C: An evaluation of elementary-reaction based and affinity-based rate laws. Geochim. Cosmoehim. Acta 59, 1457-1471. Simkiss K. (1964) Variation in the crystallization form of calcium carbonate from artificial sea water. Nature 201, 492-493. Sj6berg E. L. and Rickard D. T. (1984a) Calcite dissolution kinetics: Surface speciation and the origin of the variable pH dependence. Chem. Geol. 42, 119-136. Sj6berg E. L. and Rickard D. T. (1984b) Temperature dependence of calcite dissolution kinetics between 1 and 62C at pH 2.7 to 8.4 in aqueous solution. Geochim. Cosmochim. Acta 48, 485493. Speer J. A. (1983) Crystal chemistry and phase relations of orthorhombic carbonates. Rev. Mineral. MSA Short Course 11, 145190. Taft W. H. (1967) Physical chemistry of formation of carbonates. In Developments in Sedimentology, 9b Carbonate Rock (ed. G. V. Chilingar et al.). pp. 151-168. Elsevier. Talman S. J., Wiwchar B., Gunter W. D., and Scarge C. M. (1990) Dissolution kinetics of calcite in the H20-CO2 system along the steam saturation curve to 210C. In Fluid-Mineral Interaction (ed. R. J. Spencer and I. M. Chou); Geochem. Soe. Spec. 2, 41-55. Taylor T. R. (1989) The influence of calcite dissolution on reservoir porosity in Miocene sandstones, Picaroon field, offshore Texas Gulf Coast. J. Sediment. Petrol. 60, 322-334. Weedman S. D., Brantley S. L., and Albrecht W. (1992) Secondary compaction after secondary porosity: Can it form a pressure seal? Geology 20, 303-306.

You might also like